Skip to main content
Pharmacological Reviews logoLink to Pharmacological Reviews
. 2012 Apr;64(2):359–388. doi: 10.1124/pr.111.004697

Serotonin and Blood Pressure Regulation

Stephanie W Watts 1,, Shaun F Morrison 1, Robert Patrick Davis 1, Susan M Barman 1
Editor: Lynette C Daws
PMCID: PMC3310484  PMID: 22407614

Abstract

5-Hydroxytryptamine (5-HT; serotonin) was discovered more than 60 years ago as a substance isolated from blood. The neural effects of 5-HT have been well investigated and understood, thanks in part to the pharmacological tools available to dissect the serotonergic system and the development of the frequently prescribed selective serotonin-reuptake inhibitors. By contrast, our understanding of the role of 5-HT in the control and modification of blood pressure pales in comparison. Here we focus on the role of 5-HT in systemic blood pressure control. This review provides an in-depth study of the function and pharmacology of 5-HT in those tissues that can modify blood pressure (blood, vasculature, heart, adrenal gland, kidney, brain), with a focus on the autonomic nervous system that includes mechanisms of action and pharmacology of 5-HT within each system. We compare the change in blood pressure produced in different species by short- and long-term administration of 5-HT or selective serotonin receptor agonists. To further our understanding of the mechanisms through which 5-HT modifies blood pressure, we also describe the blood pressure effects of commonly used drugs that modify the actions of 5-HT. The pharmacology and physiological actions of 5-HT in modifying blood pressure are important, given its involvement in circulatory shock, orthostatic hypotension, serotonin syndrome and hypertension.

I. Introduction

5-Hydroxytryptamine (5-HT;1 serotonin) is an ancient substance (Azmitia, 2001). The discovery of 5-HT is part of pharmacological history. 5-HT was recognized as a substance, isolated from blood serum (sero-), that could modify the tone of smooth muscle (-tonin) (Rapport et al., 1948; Erspamer and Asero, 1952; Page and McCubbin, 1953a,b). Just a few years later, the two original 5-HT receptors—D for dibenzyline and M for morphine—were recognized in smooth muscle preparations by Gaddum and Picarelli (1957). 5-HT pharmacology was born. Although 5-HT was discovered within the cardiovascular (CV) system, it is fair to say that the effects of 5-HT within the cardiovascular system are not well understood and integrated compared with the well established actions of 5-HT in the gastrointestinal system, and the plethora of knowledge regarding the actions of 5-HT in the central nervous system (Barnes and Sharp, 1999; Hoyer et al., 2002; Green, 2006; Berger et al., 2009).

This review represents an unbiased presentation of 5-HT as a substance that can modify blood pressure. We refer the reader to other reviews that cover different aspects of the CV system or that provide a more detailed historical perspective of 5-HT in the CV system: Kuhn et al., 1980; Marwood and Stokes, 1984; Docherty, 1988; Vanhoutte, 1991; van Zwieten et al., 1992; McCall and Clement, 1994; Yildiz et al., 1998; Nebigil and Maroteaux, 2001; Ramage, 2001; Doggrell, 2003; Côté et al., 2004; Maurer-Spurej, 2005; Watts, 2005; Villalón and Centurión, 2007; Ramage and Villalón, 2008; Nalivaiko and Sgoifo, 2009; Nichols, 2009; Nigmatullina et al., 2009; Monassier et al., 2010; and Mercado et al., 2011. We will not discuss pulmonary blood pressure or pulmonary hypertension, but refer readers to an excellent review: MacLean and Dempsie, 2009.

II. 5-Hydroxytryptamine Biochemistry and Models

5-HT synthesis begins with dietary intake of l-tryptophan, an essential amino acid (Fig. 1). Foods high in l-tryptophan include egg whites, cod, chocolate, dairy products (yogurt, cheeses, milk), several meats, and nuts. The fate of tryptophan lies in the comparative activities of the enzymes indoleamine dioxygenase (IDO)/tryptophan dioxygenase and tryptophan hydroxylase (TPH). A majority of l-tryptophan is handled by IDO/tryptophan dioxygenase, an estimated 5 to 10% of tryptophan being shuttled through the TPH/5-HT pathway (Salter et al., 1995; Stone and Darlington, 2002). Over the past decade, the field has recognized two independent forms of TPH. TPH1 is expressed primarily in peripheral tissues, whereas TPH2 is expressed primarily in the central nervous system (Walther and Bader, 2003; Walther et al., 2003). Splice variants of TPH2 have been observed (Abumaria et al., 2008). This enzyme, dependent on the important cofactor tetrahydrobiopterin, commits tryptophan to the fate of 5-HT by converting tryptophan to 5-hydroxytryptophan (5-HTP; Fig. 1) (Kuhn 1999). Mouse knockouts of both the TPH1 and TPH2 (Alenina et al., 2009) isoform are available, as is a double knockout of TPH1 and TPH2 (Savelieva et al., 2008). A multitude of aromatic amino acid decarboxylases can then convert 5-HTP into 5-HT. 5-HT is rapidly converted by monoamine oxidase and aldehyde dehydrogenase to 5-hydroxyindole acetic acid (5-HIAA), a stable metabolite. 5-HT itself can also be converted into melatonin (Stone and Darlington, 2002), whereas the IDO product, kynurenine, has niacin as one of its downstream products. Thus, ingestion of tryptophan is not a pure commitment to 5-HT synthesis. This is a relevant statement, given that tryptophan is sold as a dietary supplement, in an unregulated fashion, intended for use an antidepressant and sleep aid. The molecules depicted in Fig. 1 and those in the tryptophan metabolic pathway can be identified with high-pressure liquid chromatography.

Fig. 1.

Fig. 1.

Chemical schematic of the synthesis and metabolism of 5-HT.

In many cells, 5-HT is taken up by the serotonin transporter (SERT) and by other amine transporters. Both mouse and rat models with a nonfunctional SERT have been produced and will be further discussed in section IV.A. In each species, circulating blood 5-HT is low, but intestinal 5-HT concentration is not different in wild-type and the SERT(−/−) animals (Chen et al., 2001; Linder et al., 2009). Thus, the ability of the body to make 5-HT is not compromised, but the ability of the platelet to carry 5-HT is significantly reduced when SERT is not functional. An outstanding question is the fate of intestinal 5-HT without the platelet. The SERT dysfunctional rodents, along with the TPH knockouts, have facilitated important in vivo investigations of the relationship between 5-HT and blood pressure, as described in succeeding sections.

III. 5-Hydroxytryptamine Receptor Pharmacology

As described above, 5-HT receptor pharmacology began with the discovery of the D receptor (dibenzyline) and M receptor (morphine) in the guinea pig ileum (Gaddum and Picarelli, 1957). The International Union of Basic and Clinical Pharmacology (IUPHAR) issues receptor nomenclature guidelines, and this was last done for 5-HT in publication form in 1994 (Hoyer et al., 1994; Martin, 1994). IUPHAR routinely updates its database; thus, the most current information, as supported by experts in 5-HT, can be found at this site: http://www.iuphar-db.org/DATABASE/FamilyMenuForward?familyId=1. This particular site covers all 5-HT receptors except for the 5-HT3 receptor, which can be found at http://www.iuphar-db.org/DATABASE/FamilyMenuForward?familyId=68. This division recognizes that most 5-HT receptors are G protein-coupled, heptahelical proteins (class A), whereas the 5-HT3 receptors are ion channels. In Table 1, we present a straightforward nomenclature scheme for 5-HT receptors. Many of the responses to 5-HT receptor activation cited in this table will be described below, and it is clear that 5-HT has significant and diverse effects throughout the cardiovascular system.

TABLE 1.

5-HT receptor pharmacology and location in the CV system

Values in parenthese are Ki values. Cloned indicates that values are only from cloned, transfected receptors. Affinity values are from averaged experiments available on http://pdsp.med.unc.edu/indexR.html. Modified from Watts, 2005.

Receptor Agonists Antagonists Location and Response Relevant to Blood Pressure
Heptahelical graphic file with name zpg002122346009a.jpg
Ion channel graphic file with name zpg002122346009b.jpg
5-HT1A (2.65 nM) graphic file with name zpg002122346009c.jpg 8-OH-DPAT, U92016A WAY100635, NAN190 Central nervous system (lower, raise blood pressure)
5-HT1B (16.01 nM) graphic file with name zpg002122346009d.jpg CP-93129, sumatriptan, eletriptan (some 1D affinity) GR-127935 (some 1D affinity), GR55562, isamoltane, SB236057 Smooth muscle (contraction); sympathetic presynaptic terminal (inhibition of NE release); sympathetic ganglia (inhibit transmission); central nervous system (lower, raise blood pressure)
5-HT1D (10.05 nM) graphic file with name zpg002122346009e.jpg PNU-109291, alniditan, eletriptan (some 1B affinity), L-703,664 SB 272183, LY310762, BRL15572 Smooth muscle (contraction)
5-ht1Ea (7.00 nM, cloned) graphic file with name zpg002122346009f.jpg 5-CT (nonselective), BRL54443 (some 1F affinity) None available None identified
5-HT1F (67.60 nM, cloned) graphic file with name zpg002122346009g.jpg LY334370, BRL54443 (1E affinity), LY344864 None available Smooth muscle (contraction), trigeminal nerve
5-HT2A (970.80 nM) graphic file with name zpg002122346009h.jpg DOB (2A/2B, 2C), DOI, α-methyl-5-HT (non-selective), TCB-2 R-96544, MDL100907, volinanserin Platelet (aggregation and 5-HT release); smooth muscle (contraction); adrenal gland (epinephrine release); heart (tachycardia, contraction); central nervous system
5-HT2B (11.35 nM) graphic file with name zpg002122346009i.jpg BW723C86 LY272015, RS127445 Endothelium (relaxation); smooth muscle (contraction); cardiac valves (proliferation)
5-HT2C (32.58 nM) graphic file with name zpg002122346009j.jpg WAY 163909, DOI (2A, 2B, 2C), MK212, 1-methylpsilocin RS102221, SB242084 None identified
5-HT3 (splice variants) (190.33 nM) graphic file with name zpg002122346009k.jpg SR57227A, 2-methyl-5-HT, phenylbiguanide Odansetron, granisetron Vagus nerve (bradycardia); ganglia
5-HT4(short,long) (117.00 nM, cloned) graphic file with name zpg002122346009l.jpg RS67506, BIMU1, BIMU8, zacopride GR113808, RS100235, SB204070 Cardiomyocyte (contraction)
5-HT6 (116.53 nM, cloned) graphic file with name zpg002122346009n.jpg WAY 181,187, EMD 386088 Ro 04-6790, SB 399885, SB271046 None identified
5-HT7(a–d) (3.65 nM, cloned) graphic file with name zpg002122346009o.jpg LP12, LP44, AS-19, 5-CT (nonselective) SB-269970, SB-258719 Smooth muscle (relaxation); cardiomyocyte (contraction)
*

No physiological response has been associated with receptor, hence lower case.

AS-19, (2S)-(+)-5-(1,3,5-trimethylpyrazol-4-yl)-2-(dimethylamino)tetralin; BIMU1, endo-N-(8-methyl-8-azabicyclo-[3.2.1] oct-3-yl)-2,3-dihydro-3-ethyl-2-oxo-1H-benzimidazole-1-carboxamide; BIMU8, 2,3-dihydro-N-[(3-endo)-8-methyl-8-azabicyclo[3.2.1]oct-3-yl]-3-(1-methylethyl)-2-oxo-1H-benzimidazole-1-carboxamide; BRL15572, 3-[4-(4-chlorophenyl)piperazin-1-yl]-1,1-diphenyl-2-propanol; BRL54443, 5-hydroxy-3-(1-methylpiperidin-4-yl)-1H-indole; BW723C86, α-methyl-5-(2-thienylmethoxy)-1H-indole-3-ethanamine; CP-93129, 1,4-dihydro-3-(1,2,3,6-tetrahydro-4-pyridinyl-5H-pyrrol[3,2-b]pyridin-5-one; DOI, 2,5-dimethoxy-4-iodoamphetamine; EMD 386088, 5-chloro-2-methyl-3-(1,2,3,6-tetrahydro-4-pyridinyl)-1H-indole; GR113808, 1-methyl-1H-indole-3-carboxylic acid, [1-[2-[(methylsulfonyl)amino]ethyl]-4-piperidinyl]methyl ester; GR-127935, N-[4-methoxy-3-(4-methyl-1-piperazinyl)phenyl]-2′-methyl-4′-(5-methyl-1,2,4-oxadiazol-3-yl)-1,1′-biphenyl-4-carboxamide; GR55562, 3-[3-(dimethylamino)propyl]-4-hydroxy-N-[4-(4-pyridinyl)phenyl]benzamide; L-703,664, N,N-dimethyl-5-[(5-methyl-1,1-dioxodo-1,2,5-thiadiazolidin-2-yl)methyl]-1H-indole-3-ethanamine; LP12, 4-(2-diphenyl)-N-(1,2,3,4-tetrahydronaphthalen-1-yl)-1-piperazinehexanamide; LP44, 4-[2-(methylthio)phenyl]-N-(1,2,3,4-tetrahydro-1-naphthalenyl)-1-piperazinehexanamide; LY272015, 1-[(3,4-dimethoxyphenyl)methy]-2,3,4,9-tetrahydro-6-methyl-1H-pyrido[3,4-b]indole; LY310762, 1-[2-[4-(4-fluorobenzoyl)-1-piperidinyl]ethyl]-1,3-dihydro-3,3-dimethyl-2H-indol-2-one; LY334370, 4-fluoro-N-(1-methyl-4-piperidinyl)-1H-indol-5-yl]-benzamide; LY344864, N-[(3R)-3-(dimethylamino)-2,3,4,9-tetrahydro-1H-carbazol-6-yl]-4-fluorobenzamide; MDL100907, (2,3-dimethoxyphenyl)-[1-[2-(4-fluorophenyl)ethyl]piperidin-4-yl]methanol; MK212, 6-chloro-2-(1-piperazinyl)pyrazine; NAN190, 1-(2-methoxyphenyl)-4-(4-phthalimidobutyl)piperazine; PNU-109291, (S)-3,4-dihydro-1-[2-[4-(4-methoxyphenyl)-1-piperazinyl]ethyl]-N-methyl-1H-2-benzopyran-6-carboxamide; R-96544, (2R,4R)-5-[2-[2-[2-(3-methoxyphenyl)ethyl]phenoxy]ethyl]-1-methyl-3-pyrrolidinol; Ro 04-6790, 4-amino-N-[2,6-bis(methylamino)-4-pyrimidinyl]-benzenesulfonamide; RS100235, 1-(5-amino-6-chloro-2,3-dihydro-1,4-benzodioxin-8-yl)-3-[1-[3-(3,4-dimethoxyphenyl)propyl]piperidin-4-yl]propan-1-one; RS102221, 8-[5-(2, 4-dimethoxy-5-(4-trifluoromethylphenylsulphonamido)phenyl-5-oxopentyl]-1,3,8-triazaspiro[4.5]decane-2,4-dione; RS127445, 4-(4-fluoro-1-naphthalenyl)-6-(1-methylethyl)-2-pyrimidinamine; RS67506, 1-(4-amino-5-chloro-2-methoxyphenyl)-3-[1–2-methylsulphonylamino)ethyl-4-piperidinyl]-1-propanone; SB204070, (1-butyl-4-piperidinyl)methyl-8-amino-7-chloro-1,4-benzodioxane-5-carboxylate; SB236057, 1′-ethyl-7-({4-[2-methyl-4-(5-methyl-1,3,4-oxadiazol-2-yl)phenyl]phenyl}carbonyl)-2,5,6,7-tetrahydrospiro[furo [2,3-f]indole-1,4′-piperidine]; SB242084, 6-chloro-2,3-dihydro-5-methyl-N-[6-[(2-methyl-3-pyridinyl)oxy]-3-pyridinyl]-1H-indole-1-carboxyamide dihydrochloride; SB258719, 3-methyl-N-[(1R)-1-methyl-3-(4-methyl-1-piperidinyl)propyl]-N-methylbenzenesulfonamide; SB269970, (2R)-1-[(3-hydroxyphenyl)sulfonyl]-2-[2-(4-methyl-1-piperidinyl)ethyl]pyrrolidine; SB271046, 5-chloro-N-[4-methoxy-3-(1-piperazinyl)phenyl]-3-methyl-benzo[b]thiophen-2-sulfonamide; SB272183, 5-chloro-2,3-dihydro-6-(4-methylpiperazin-1-yl)-1(4-pyridin-4-yl)napth-1-ylaminocarbonyl]-1H-indole); SB399885, N-(3,5-dichloro-2-methoxyphenyl)-4-methoxy-3-(1-piperazinyl)benzenesulfonamide; SB699551, N-[2-(dimethylamino)ethyl]-N-[[4′-[[(2-phenylethyl)amino]methyl][1,1′-biphenyl]-4-yl]methyl]cyclopentanepropanamide; SR57227A, 1-(6-chloro-2-pyridinyl)-4-piperidinamine; U92016A, (R)-8-dipropylamino-6,7,8,9-tetrahydro-3H-benzo[e]indole-2-carbonitrile; WAY 100635, N-[2-[4-(2-methoxyphenyl)-1-piperazinyl]ethyl]-N-(2-pyridyl)cyclohexanecarboxamide; WAY 163909, [(7bR, 10aR)-1,2,3,4,8,9,10, 10a-octahydro-7bH-cyclopenta-[b][1,4]diazepino[6,7,1-hi]indole]; WAY 181,187, 2-(1-{6-chloroimidazo[2,1-b][1,3]thiazole-5-sulfonyl}-1H-indol-3-yl)ethan-1-amine.

IV. Intersections of 5-Hydroxytryptamine and the Cardiovascular System that Affect Blood Pressure

Several tissues, organs, and neural circuits contribute to the control of blood pressure, and 5-HT can influence many of them. Table 1 lists the relevant locations of the various 5-HT receptors that contribute to cardiovascular regulation. Although we will briefly describe 5-HT's effects in the kidney, adrenal gland, heart, and blood, the focus of this review is on the vasculature, its control by the sympathetic nervous system, and the central nervous system pathways that determine the sympathetic nerve activity to cardiovascular tissues. We describe the presence and handling of 5-HT, the function and pharmacology of 5-HT, and give at least one example of a change in the response to 5-HT in a pathological condition that is specific to each system.

A. 5-Hydroxytryptamine in Circulating Blood: Free and Platelet-Bound

It is here that the highest peripheral levels of 5-HT are found in the cardiovascular system. Understanding the handling of 5-HT (synthesis, storage, release) in blood is critical, because it is through the blood that the circulatory system will come into contact with free, circulating 5-HT as well as 5-HT contained in platelets.

1. 5-Hydroxytryptamine Synthesis and Handling.

Blood platelets do not synthesize 5-HT, but possess the SERT and acquire a high concentration of 5-HT (estimated in the millimolar range) from the intestine, where 5-HT is synthesized in enterochromaffin cells (Berger et al., 2009). Biologically active 5-HT (i.e., in contact with vasculature) is free 5-HT, existing outside of the platelet and measured as platelet free/poor 5-HT (see Table 2 for free and platelet 5-HT in humans). Rat free plasma 5-HT has been measured in the low to mid nanogram per milliliter range, largely consistent with those of the human. One to ∼100 nM concentrations of 5-HT are estimated as platelet free plasma in the human (Kema et al., 2001; Monaghan et al., 2009; estimating a 5-HT molecular weight of 176 g/mol), although Brand and Anderson (2011) question the validity of 5-HT plasma measures in humans given the high variability of 5-HT concentration reported in 101 studies that they compare. The existence of whole-blood monoamine oxidase activity implies that blood, like tissues, has the ability to metabolize 5-HT to a less active substance (Celeda and Artigas, 1993) and keep circulating levels of 5-HT low. The SERT(−/−) rat, created by Edwin Cuppen (Homberg et al., 2007), provides a unique view into other ways for 5-HT to be carried in blood. The SERT protein is truncated in this knockout such that the SERT protein, which is partially expressed, is not functional. This animal has low circulating platelet-poor and platelet-rich levels of 5-HT compared with the SERT (+/+) rat, which is to be expected because SERT is thought to be the primary mechanism by which 5-HT is concentrated in platelets (Linder et al., 2008a,b). However, when SERT (−/−) rats are infused with 5-HT, platelets do take up 5-HT (Davis et al., 2011). This suggests either incomplete knockout of SERT or, more likely, the existence of nonSERT mechanisms to concentrate 5-HT. The SERT KO mouse has a similar reduction in blood 5-HT (Chen et al., 2001). Whether there is sufficient free 5-HT in the blood to activate vascular 5-HT receptors depends on the 5-HT receptor expressed by a blood vessel, because the affinity of 5-HT for these receptors can vary significantly (Table 1). Circulating blood is the most immediate source of 5-HT for the peripheral circulatory system.

TABLE 2.

Blood 5-HT measures (whole, platelet-rich, and platelet poor) in humans

Measures of 5-HT Comments Reference
95–116 ng/ml Whole blood Jelen et al., 1979
378–518 ng/109 platelets Whole platelet Topsakal et al, 2009
65–250 ng/ml Whole blood Breuer et al., 1996
1–3 ng/ml Blood diasylate Castejon et al., 1999
0–609 ng/109 platelets Platelet-rich plasma Koch et al., 2004
350–650 ng/109 platelets Platelet-rich plasma (female) Carrasco et al., 1998
1.0–1.6 nM Platelet rich plasma Brenner et al., 2007
3.4–23.8 nmol/109 platelets Platelet rich plasma Kema et al., 2001
66–106 ng/108 platelets Platelet rich plasma Biondi et al., 1988
105–165 nmol/1011 platelets Platelets Kamal et al., 1984
88–246 ng/108 platelets Platelets Gujrati et al., 1994
3.4–3.5 nmol/109 platelets Platelets Fetkovska et al., 1990
12–20 ng/ml Platelet-poor plasma (children) Breuer et al., 1996
1–4 ng/ml Platelet-poor plasma (female) Carrasco et al., 1998
0.1–1 nM Platelet-poor plasma Brenner et al., 2007
2.5–6.1 ng/ml Platelet-poor plasma Biondi et al., 1988
89.5–115.5 nM Platelet-poor plasma Monaghan et al., 2009
5.6–23.9 ng/ml plasma Platelet-poor plasma Koch et al., 2004

2. Function and 5-Hydroxytryptamine Pharmacology.

The 5-HT2A receptor is the dominant receptor in platelets of multiple species. Once stimulated, 5-HT2A receptors promote the aggregation of platelets, resulting in release of more 5-HT, ADP, and other substances. Preparations of platelets cause either contraction or relaxation of isolated arteries (Zellers et al., 1991).

3. Change in Pathological Conditions.

Platelets carry most of the blood 5-HT, so changes in platelet mechanics, fragility, and aggregation (promoted by 5-HT) can dramatically change local 5-HT concentration and, potentially, circulating 5-HT (Le Quan Sang et al., 1991; Ding et al., 1994). In 1989, Umegaki et al. demonstrated reduced content of platelet 5-HT in deoxycorticosterone salt (DOCA salt) hypertensive rats. This is a finding consistent with a number of models of hypertension and suggests that, at least in hypertension, the platelet is “activated,” in that it has a lower content of 5-HT.

B. 5-Hydroxytryptamine Presence and Function in the Vasculature

5-HT was discovered, in part, as a vasoconstrictor and this is the property for which 5-HT is best known in the cardiovascular system.

1. 5-Hydroxytryptamine Synthesis and Handling.

A serotonergic system (uptake, synthesis, and metabolism) exists in isolated blood vessels (Ni et al., 2008). Isolated blood vessels from the mouse and rat stain for antibodies raised against 5-HT, a staining that does not depend on the presence of resident circulatory cells that might contain 5-HT. Ni et al. (2004) demonstrated the presence of TPH1 but not TPH2 mRNA and protein in the isolated normal rat aorta and superior mesenteric artery, as well as the ability of the isolated artery to synthesize 5-HTP when arteries were incubated with exogenous tryptophan and BH4. Isolated blood vessels—including arteries and veins—concentrate 5-HT actively through SERT (Ni et al., 2004; Linder et al., 2008a,b). Fenfluramine can release 5-HT from the blood vessel into the fluid in which the vessel is bathed (Ni et al., 2008), and this 5-HT would be near 5-HT receptors that exist within the blood vessel. Thus, there are at least two sources of 5-HT for a blood vessel: endogenously synthesized 5-HT and exogenous 5-HT. It is noteworthy that blood vessels can rapidly metabolize transported 5-HT to 5-HIAA.

2. Function and 5-Hydroxytryptamine Pharmacology.

In humans and in animals, 5-HT predominantly causes direct arterial constriction, and the list of references here are just a few that report the effects of 5-HT in the isolated artery; more are included in the legend to Fig. 2: McGregor and Smirk, 1970; Docherty, 1988; Rosón et al., 1990; Vanhoutte, 1991; van Zwieten et al., 1992; Webb et al., 1992; Nishimura and Suzuki, 1995; Nishimura, 1996; Yildiz et al., 1996, 1998; Boston and Hodgson, 1997; Watts, 1997, 2002, 2005, 2009; Hutri-Kähönen et al., 1999; Ramage, 2001; Janiak et al., 2002; Doggrell, 2003; Gul et al., 2003; Côté et al., 2004; Watts and Thompson, 2004; Villalón and Centurión, 2007; Ramage and Villalón, 2008; Nichols, 2009; and Nigmatullina et al., 2009. These studies have been performed in a variety of isolated blood vessels—basilar, superior mesenteric, aortic, femoral, mesenteric resistance, for example—such that constriction cannot be attributed to a single receptor type or size of vessel. The diversity of vessel response (C = contraction, R = relaxation), species response within a vessel and 5-HT receptor subtype mediating the response (e.g., 1B, 2A) is illustrated in Fig. 2. This figure was constructed using only data from isolated vessel studies. Cerebral vessels include the meningeal, temporal and occipital arteries. Resistance vessels are included in several of the circulations (cerebral and mesenteric). Vasoconstriction is predominantly mediated by the 5-HT2A receptor, but 5-HT1B/1D receptors can also mediate constriction, exemplified by the success of the triptans, 5-HT1B/1D agonists, in the treatment of migraine (Gilmore and Michael, 2011). Virtually every blood vessel, when mounted in a tissue bath, responds to 5-HT with contraction from baseline.

Fig. 2.

Fig. 2.

Response of the vasculature to 5-HT, as depicted by using the rat vasculature as a model. Both constriction (C) and relaxation (R) may be listed for the same species if both responses were observed. Species name is listed, and the subtype of the receptor mediating the response is listed second (1B, 2A). UK, unknown receptor mechanism. Data were collected from the following references: Lemberger et al., 1984; Miller et al., 1984; Feniuk et al., 1985; Nyborg and Mikkelsen, 1985; Cohen, 1986; Leff et al., 1987; Hamel et al., 1989, 1993; Parsons et al., 1989, 1992; Bodelsson et al., 1990; Borton et al., 1990; Chester et al., 1990; Gaw et al., 1990; Mylecharane, 1990; Toda and Okamura, 1990; van Heuven-Nolsen et al., 1990; Asher et al., 1991; Lai et al., 1991; Parsons, 1991; Sumner, 1991; Cushing and Cohen, 1992a,b, 1993; Dorigo et al., 1992; Eglen et al., 1992; Weiner et al., 1992; Bax et al., 1993; Glusa and Müller-Schweinitzer, 1993; Jansen et al., 1993; Cushing et al., 1994, 1996; Yildiz and Tuncer, 1994; Fujiwara and Chiba, 1995; Miranda et al., 1995; Schmuck et al., 1996; Terrón, 1996; Valentin et al., 1996; Verheggen et al., 1996, 1998, 2004, 2006; Zwaveling et al., 1996; Ellwood and Curtis, 1997; Parsons et al., 1998; Nilsson et al., 1999; Roon et al., 1999; Terrón and Falcón-Neri, 1999; Bouchelet et al., 2000; Galzin et al., 2000; Geerts et al., 2000; Ishida et al., 2001; Lamping and Faraci, 2001; McKune and Watts, 2001; Schöning et al., 2001; Razzaque et al., 2002; Teng et al., 2002; van den Broek et al., 2002; Froldi et al., 2003, 2008; Edvinsson et al., 2005; Nagai et al., 2007; Zerpa et al., 2007; Masu et al., 2008; Linder et al., 2010; Radenkoviä et al., 2010.

By contrast, not all blood vessels can relax to 5-HT. In the rat jugular vein and pulmonary artery and coronary arteries of rat and greyhound, 5-HT causes relaxation through activation of 5-HT2B and 5-HT7 receptors (Mylecharane, 1990; Mankad et al., 1991; Woodman and Dusting, 1994; Ellis et al., 1995; Glusa and Roos, 1996; Centurión et al., 2000, 2004; Jähnichen et al., 2005). 5-HT relaxes human umbilical arteries (Haugen et al., 1997) and dilates skeletal muscle arterioles, an important finding with respect to blood pressure (Alsip and Harris, 1992; Alsip et al., 1996). In some cases, 5-HT receptors that mediate constriction must be blocked before unmasking or revealing 5-HT relaxant receptors (McLennan and Taylor, 1984). This raises the question of how these relaxant 5-HT receptors would be activated physiologically. Do they dampen the overall effect of 5-HT vasoconstriction? In what situation(s) can relaxant receptors be directly activated without activation of contractile receptors? In an isolated tissue bath, we may not be appropriately mimicking the group of substances that would act as the endogenous constrictors upon which 5-HT would stimulate relaxation. There are examples in which 5-HT relaxant receptors are revealed without blockade of contractile 5-HT receptors. This includes the equine coronary artery (Obi et al., 1994) and dog coronary artery (Terrón, 1996). Although no antagonists of contractile 5-HT receptors were added in the equine coronary artery studies, Terrón (1996) showed that whereas 5-HT relaxation was present naturally in the dog coronary artery, the sensitivity to 5-HT and 5-carboxamidotryptamine (5-CT) as relaxants was increased when the 5-HT1B/1D receptor antagonist N-[4-methoxy-3-(4-methyl-1-piperazinyl)phenyl]-2′-methyl-4′-(5-methyl-1,2,4-oxadiazol-3-yl)-1–1′-biphenyl-4-carboxamide (GR127935) was added. We note that 5-CT has affinity for 5-HT7 receptors (Waeber and Moskowitz, 1995; Krobert et al., 2001). Thus, both contractile and relaxant receptors are in play in blood vessels, and the balance of their effects is different in different vessel types and species.

The effects of 5-HT in the vasculature become less clear when studying a system more complicated than an isolated vessel, probably because 5-HT now has the ability to stimulate multiple receptors within multiple tissue types that may act in seemingly contradictory fashion as it pertains to smooth muscle tone. We share here but a few examples. Calama et al. (2003, 2005) raise the interesting possibility that β2 adrenergic receptors are the effectors of 5-HT-induced vasodilation in the rat hindquarters, connecting 5-HT to the adrenal medulla (epinephrine release?), because 5-HT itself does not have appreciable affinity for β-adrenergic receptors. Whether released epinephrine would also interact with α-adrenergic receptors in this situation is unknown. It is interesting to note that 5-HT itself has affinity for α-adrenergic receptors (Grandaw and Purdy, 1996). In pentobarbital-anesthetized dogs, intra-arterial 5-HT caused a vasodilation in the femoral arterial circulation that was abolished by ganglionic blockade (Phillips et al., 1985). Likewise, 5-HT increases the external carotid blood flow of the dog (Villalón et al., 1993). In the human forearm vasculature, 5-HT causes an increase in forearm blood flow, but the receptor mechanism is unclear. Blauw et al. (1988, 1991), Bruning et al. (1993, 1994), and Kemme et al. (2000) performed a number of studies to identify this receptor, and the pharmacology of the 5-HT relaxant receptor was most consistent with that of the 5-HT3 receptor, a receptor unusual to the vasculature. Because of the inherent difficulty in performing studies in humans, the breadth of the pharmacology used to verify and identify this response has not been as wide as that in animal models.

The ability of 5-HT to alter arteriovenous anastomoses (AVA)—present in the skin, fingers, lips, and ears—has been studied. Saxena and Verdouw (1982) reported increases in arteriolar blood flow in pig ears and skin, observed as a “pinking” of the tissue after injection of 5-HT (1–10 μg · kg−1 · min−1) into the carotid artery of the pig, an event that decreased the blood pressure of the animals. This increased flow, presumably in the capillaries or nutritive vessels, was considered to occur at the expense of non-nutrient or AVA blood flow. Experiments were also performed in the cat, which similarly responded with a decrease in blood pressure to intra-arterial 5-HT, but tissue blood flow did not change significantly. In the human, intra-arterial 5-HT decreased AVA blood flow and increased capillary skin blood flow (Blauw et al., 1991). Likewise, the 5-HT1A agonist 8-OH-DPAT decreased AVA blood flow in the anesthetized pig (Bom et al., 1989) and caused cutaneous vasoconstriction in the canine forelimb (Dobbins et al., 1983). The affinity of 8-OH-DPAT for the 5-HT7 receptor complicates the attribution of such an event to a single receptor (Wood et al., 2000; Sprouse et al., 2004). In the rat tail, 5-HT has the interesting effect of causing an increase in tail temperature—indicative of dilation in this cutaneous circulation—that was associated with a decrease in blood pressure (Key and Wigfield, 1992). Similar results were observed with the administration of the 5-HT releaser fenfluramine in the rat (Subramanian and Vollmer, 2004). Thus, a cutaneous vasodilation in the rat may play a role in the decrease in blood pressure in response to 5-HT, but the mechanisms of such dilation are not known. A discussion of the relative contribution of the cutaneous circulation to changes in systemic blood pressure is beyond the scope of this review but is an interesting idea in light of the discussion of whether temperature homeostasis or blood pressure homeostasis has primacy in the body (Blankfield, 2006).

3. Change in Pathological Conditions.

Hyper-reactivity (or enhanced vasoconstriction) to 5-HT is a hallmark of vascular damage. This is observed experimentally as a lower threshold for 5-HT to cause contraction, an increased potency of 5-HT, and/or an increased efficacy of 5-HT compared with a normotensive control. One of the best-studied changes in vascular response to 5-HT has been in vessels from humans and animals with high blood pressure (hypertension). Hyper-reactivity occurs in arteries and in veins in hypertension, in a number of different vascular beds and differently sized vessels (Cummings et al., 1986; Thompson and Webb, 1987; Huzoor-Akbar et al., 1989; Dohi and Lüscher, 1991; Webb et al., 1992; Moreno et al., 1996). In several models of experimental hypertension, up-regulation of 5-HT2B receptors in the smooth muscle is one mechanism by which 5-HT becomes hyper-reactive (Watts et al., 1995, 1996; Watts, 1997; Watts and Fink, 1999; Banes and Watts, 2002, 2003; Russell et al., 2002). Mineralocorticoids can directly stimulate expression of the 5-HT2B receptor (Banes and Watts, 2002, 2003) such that an up-regulated receptor enables a more sensitive contraction. It should be noted that the 5-HT2B receptor is expressed in arterial smooth muscle of a normotensive animal. This receptor does not seem to be functionally coupled to contraction as it is in hypertension. Thus, there may be more than one event in hypertension—up-regulation of the receptor and/or signaling element—that allows for vessels to be hyperresponsive to 5-HT. This is but one example of the change in vascular responsiveness to 5-HT in a pathological state.

C. 5-Hydroxytryptamine Presence and Function in the Heart

The effects of 5-HT on the heart are complex, and important work has been performed, in particular by Kaumann and Levy (2006) and Nebigil and Maroteaux (2001).

1. 5-Hydroxytryptamine Synthesis and Handling.

5-HT has been found in the heart (Sole et al., 1979), and Ikeda et al. (2005) demonstrated the ability of neonatal rat cardiomyocytes to synthesize 5-HT. 5-HT is a survival factor for cardiomyocytes, as demonstrated by Nebigil et al. (2003a). SERT expression has also been observed in cultured cardiac myocytes (Sari and Zhou, 2003) and heart valves (Pavone et al., 2008). Loss of TPH in the mouse resulted in abnormal cardiac activity, suggesting that peripheral 5-HT—from within or outside of the heart—is important to cardiac function (Côté et al., 2003) and development.

2. Function and 5-Hydroxytryptamine Receptor Pharmacology.

Kaumann and Levy (2006) have provided a focus for 5-HT receptors in the human cardiovascular system, demonstrating that 5-HT increases atrial function and arrhythmias as well as positive inotropic, lusitropic, and arrhythmic effects in the ventricle, primarily through activation of 5-HT4 receptors. The work of Nebigil and Maroteaux (2001) strongly supports the role of the 5-HT2B receptor (originally the stomach fundus 5-HT receptor) in normal heart development and function. These researchers and others have shown that receptors for 5-HT exist directly on cardiac myocytes and on the vagus and sympathetic nerves. Stimulation of 5-HT3 receptors on the vagus nerves (cardiac vagal afferents) accounts for the bradycardia elicited by activation of the Bezold-Jarisch reflex. 5-HT can also act as a sympatholytic through activation of 5-HT1 receptors on sympathetic nerve terminals, inhibiting norepinephrine release. Both of these mechanisms contribute to a reduced cardiac output, which would be associated with a decrease in blood pressure. However, 5-HT can also activate the heart. The receptor mechanism for 5-HT to increase heart rate (positive chronotropy) is species-dependent. For example, this occurs through activation of 5-HT2A receptors in the rat, 5-HT4 receptors in the pig and human, and 5-HT7 receptors in the cat (Saxena and Villalón, 1991; Villalón et al., 1997; Côté et al., 2004). In isolated cardiomyocytes, 5-HT stimulates mitogenesis, and the 5-HT2B receptor is critical for development of the heart in the mouse (Nebigil and Maroteaux, 2001). 5-HT stimulates hypertrophy in cardiac myocytes, and Bianchi et al. (2005) suggested the provocative hypothesis that 5-HT may be used as a substrate for monoamine oxidase (MAO), ultimately providing the hydrogen peroxide that activates hypertrophic pathways. Thus, the cardiac effects of 5-HT are complex and species-dependent; collectively, these studies suggest that 5-HT could both increase and decrease blood pressure through its actions in the heart.

3. Change in Pathological Conditions.

Up-regulation and stimulation of the 5-HT2B receptor within the heart leads to cardiac hypertrophy (Nebigil et al., 2003b), such that mice lacking the 5-HT2B receptor are protected from cardiac hypertrophy. Significant interest has been paid to the expression of 5-HT receptors on aortic valves (especially the 5-HT2B receptor), because the 5-HT releaser and weight-loss drug fenfluramine causes valvular dysfunction (Roth, 2007; Huang et al., 2009; Hajjo et al., 2010). Finally, SERT KO mice demonstrate cardiac fibrosis (Pavone et al., 2009).

D. 5-Hydroxytryptamine Presence and Function in the Kidney

1. 5-Hydroxytryptamine Synthesis and Handling.

The proximal tubules of the kidney are a site of synthesis for 5-HT (Stier and Itskovitz, 1985; Sole et al., 1986; Stier et al., 1986; Itskovitz et al., 1989; Hafdi et al., 1996). As a positively charged molecule, 5-HT is excreted by SERT and other cation transporters within the nephron, where it may reduce the inhibitory effect of parathyroid hormone-induced inhibition of sodium-phosphate transport (Hafdi et al., 1996).

2. Function and 5-Hydroxytryptamine Pharmacology.

5-HT has two notable effects on renal function, both of which would promote the elevation of blood pressure. First, when administered to the isolated, perfused kidney, 5-HT causes an elevation in perfusion pressure, a response supported by renal vasoconstriction. The 5-HT receptor of the main renal artery is exquisitely sensitive to 5-HT, and the pharmacology of this interesting response has yet to be defined, most closely resembling a 5-HT2-like receptor in the rat (Watts and Thompson, 2004). Second, when 5-HT is given to rats or cats in vivo or in vitro to cortical tubules, sodium excretion is reduced (Fastier and Waal, 1957; Frandsen and Nielsen, 1966; Soares-da-Silva, 1996). In the cat (Fastier and Waal, 1957), 5-HT caused a decrease in blood pressure. In the human, the prodrug γ-l-glutamyl-5-hydroxy-l-tryptophan was given (16.6 μg · kg−1 · min−1 i.v.) over 60 min. The authors state that there was no difference in blood pressure and pulse rate on the days of experimentation, but the data are not presented (Li Kam Wa et al., 1994). Sodium and volume retention would keep blood volume elevated. In the dog, the picture is less clear, because 5-HT has been reported to increase (Shoji et al., 1989) or more commonly decrease (Blackmore, 1958; Park et al., 1968) urinary excretion of sodium. Overall, these findings suggest 5-HT has antinatriuretic/antidiuretic effects within the kidney that increase blood volume and support blood pressure.

3. Change in Pathological Condition.

Serotonin syndrome, which results from an elevated level of 5-HT typically caused by ingestion of foods and drugs, can result in renal failure (Rajapakse et al., 2010). 5-HT has also been implicated in the nephropathies that accompany diabetes (Doggrell, 2003).

E. 5-Hydroxytryptamine Presence and Function in the Adrenal Gland

The adrenal gland is a significantly understudied tissue with respect to 5-HT.

1. 5-Hydroxytryptamine Synthesis and Handling.

When animals are infused with 5-HT for 1 week via a minipump (ALZET Osmotic Pumps, Cupertino, CA), the adrenal gland accumulates a significant amount of 5-HT (Linder et al., 2009), and Fig. 3 illustrates that the adrenal gland becomes one of the most concentrated reservoirs of 5-HT in the body. The uptake of 5-HT by the adrenal gland was virtually abolished in the SERT(−/−) rat, demonstrating the importance of SERT to the adrenal accumulation of 5-HT. It is noteworthy that endogenous 5-HT has also been located in the adrenal medulla, with evidence that 5-HT is synthesized within the chromaffin cells in the frog and rat (Csaba and Sudár, 1978; Verhofstad and Jonsson, 1983; Holzwarth et al., 1984; Brownfield et al., 1985; Holzwarth and Brownfield, 1985; Delarue et al., 1992). However, it is not clear whether the adrenal gland—cortex or medulla—can synthesize 5-HT. In the human, TPH was not detected immunohistochemically in the adrenal cortex (Meyer and Brinck 1999). An interesting connection between adrenal gland and brain 5-HT synthesis was made by Miller et al. (1980), who demonstrated that brain tryptophan utilization was determined by the presence or absence of the adrenal gland. GTP cyclohydrolase is present in the adrenal gland, and this makes the tetrahydrobiopterin that is a necessary cofactor of the tyrosine, tryptophan, and phenylalanine hydroxylases (Nagatsu et al., 1995). Thus, there is potential for 5-HT to modify blood pressure through actions in the adrenal gland.

Fig. 3.

Fig. 3.

Distribution of 5-HT in organs when infused with vehicle or 5-HT at a dose of 25 μg · kg−1 · min−1 in the rat for one week. Vehicle animals receive saline. Bars represent means ± S.E.M. for n = 6. Boxes highlight organs that are of cardiovascular interest. [Adapted from Linder AE, Beggs KM, Burnett RJ, and Watts SW (2009) Body distribution of infused serotonin in rats. Clin Exp Pharmacol Physiol 36:599–601 Copyright © 2009 John Wiley & Sons, Inc. Used with permission.

2. Function and 5-Hydroxytryptamine Pharmacology.

5-HT can act as a sympathomimetic in the sympathetic nerves of the blood vessels (Kawasaki and Takasaki, 1984), and 5-HT stimulates adrenal medullary epinephrine release through mechanisms that are both receptor-dependent and -independent (Bagdy et al., 1989; Sugimoto et al., 1996). 5-HT also has effects within the adrenal cortex. The role of the 5-HT4 receptor in the normal human adrenal gland includes stimulation of cortisol (Louiset et al., 2004), whereas 5-HT7 receptors are associated with an increase in adenylate cyclase activity that leads to aldosterone production (Contesse et al., 1999; Lenglet et al., 2002), suggesting an important mechanism in water and salt retention.

3. Change in Pathological Condition.

Little is known in this regard. 5-HT4 receptors are located in the adrenal gland (Brudeli et al., 2010), and elevated levels of the transcript of this receptor are detected in the adrenal glands of patients with unilateral aldosterone-producing adenoma (Cartier et al., 2005; Ye et al., 2007), but it is not known how or whether the up-regulated 5-HT4 receptor contributes to the disease. The ability of the adrenal gland to concentrate 5-HT to levels that are 2 to 3 times that in another tissue in the body (rat) raises the question of what 5-HT does in this tissue and how this is modified in disease.

F. 5-Hydroxytryptamine Influence on the Autonomic Control of Blood Pressure

5-HT has a multitude of effects on the peripheral nervous system that can ultimately modify the effects of sympathetic activity. Figure 4 depicts various places within the nervous system where 5-HT might act to affect blood pressure, and we begin at the level of the sympathetic-vascular junction.

Fig. 4.

Fig. 4.

Schematic of the potential sites at which 5-HT could interact to lower blood pressure within the context of the sympathetic nervous system. Small geometric shapes represent individual classes of 5-HT receptors; semicircular red arrow indicates inhibition.

1. Peripheral Effects of 5-Hydroxytryptamine at the Sympathetic-Vascular Junction.

The arterial vascular system, including resistance arterioles, is innervated by the sympathetic nervous system. Sympathetic nerve terminals release the sympathetic neurotransmitters—norepinephrine (NE), neuropeptide Y, and ATP—to contract vascular smooth muscle. The norepinephrine transporter (NET) is also present on the sympathetic terminal to facilitate reuptake of NE. Neurotransmitter release is regulated locally by feedback inhibition through autoreceptors (adrenergic receptors that inhibit further NE release), and feedback inhibition of NE release can occur as a result of other chemicals, including 5-HT. Feuerstein (2008) published a comprehensive review of presynaptic receptors for dopamine, histamine, and 5-HT. Although autoinhibition of serotonergic nerves through 5-HT receptors makes intuitive sense, the finding of 5-HT receptors on adrenergic nerves is less intuitive. The 5-HT1B/1D receptor class has been found on sympathetic nerve terminals (Feniuk et al., 1979; Shepherd and Vanhoutte, 1985; Göthert et al., 1986, 1991; Molderings et al., 1990). Activation of these receptors results in a reduced NE release from the terminal such that the overall effect is reduced contractile tone.

Likewise, the presence of NET in sympathetic terminals is reasonably understood. Whether SERT has also been localized to the vascular sympathetic terminal remains in question. The promiscuity of these transporters—SERT taking up NE, NET taking up 5-HT—is recognized (Daws, 2009). This promiscuity can have physiological relevance. In a fascinating study, Kawasaki and Takasaki (1984) demonstrated that 5-HT could be taken up and released from sympathetic terminals in the vasculature. This raises the intriguing possibility that 5-HT may be coreleased with NE in normal transmission; this is especially important when recognizing that 5-HT has the ability to potentiate the contractile response of arteries to endogenous hormones such as NE (MacLennan et al., 1993; Yildiz et al., 1998). In mesenteric arteries, 5-HT-like immunoreactive nerves have been reported (Gale and Cowen, 1988). Although circulating levels of 5-HT are typically low as a result of platelet uptake of 5-HT via SERT, a thrombotic event that would cause platelet aggregation could lead to a local increase of 5-HT, such that NET would take it up and it would be either metabolized or repackaged. Whether this can happen in all sympathetic nerves remains in question.

Morán et al. (2010) demonstrated that 5-HT inhibits the pressor effect of sympathetic stimulation in long-term diabetic pithed rats. They have suggested that both the 5-HT1A and 5-HT2 receptors are involved in this response. This would implicate a variety of receptors—5-HT1A, 5-HT1B/1D, 5-HT2—in the inhibition of sympathetic activity independent of the central nervous system.

2. Peripheral Effects of 5-Hydroxytryptamine on Sympathetic Ganglionic Transmission.

a. 5-Hydroxytryptamine synthesis and handling.

The enzymes for dedicated 5-HT synthesis—tryptophan hydroxylase and an aromatic acid amino decarboxylase—exist in sympathetic ganglia. 5-HT has been immunohistochemically localized in sympathetic ganglia (Verhofstad and Jonsson, 1983; Dun et al., 1984; Häppölä, 1988; Päivärinta et al., 1989; Karhula et al., 1995). mRNA for SERT (Nishimura et al., 1999) and multiple 5-HT receptors (Newberry et al., 1996; Pierce et al., 1996; Watkins and Newberry, 1996) are present in ganglia. Removal of celiac superior mesenteric ganglion results in a loss of 5-HT-like immunoreactive nerves around mesenteric blood vessels. All of these data point to the ganglia as a site of 5-HT synthesis, uptake, and possible release. Given the integral nature of the ganglia to autonomic neurotransmission, understanding the role of 5-HT in this site is important.

b. Function and 5-hydroxytryptamine pharmacology.

Many studies using an in vitro preparation have shown that 5-HT can facilitate transmission within autonomic ganglia. Hertzler (1961) reported 5 decades ago that 5-HT decreased the threshold and increased the amplitude of the rat stellate ganglionic responses to preganglionic stimulation. Wallis and Dun (1987) demonstrated that 5-HT induced depolarization of the guinea pig celiac ganglion. Meehan and Kreulen (1991) and Cai et al. (1999) later extended this observation to include depolarization of the inferior mesenteric ganglion. Meehan and Kreulen (1991) found that this effect was dependent on activation of 5-HT3 receptors. In the superior cervical ganglion, 5-HT causes a depolarization and enhanced transmission (Watkins and Newberry 1996) by acting at more than one 5-HT receptor subtype, including 5-HT2A and 5-HT3 receptors. Thus, there seems to be an overall ability of 5-HT to stimulate ganglionic transmission, but there are exceptions to this generalization. There is evidence from a few in vitro studies to implicate an inhibitory role of 5-HT in ganglionic transmission. Gilbert and Newberry (1987) showed that 5-HT acts on 5-HT1-like receptors to hyperpolarize the cervical ganglion of the rat. Dun and Karczmar (1981) showed that 5-HT inhibited ganglionic transmission by a presynaptic mechanism (i.e., by reducing acetylcholine release from the presympathetic nerve terminal).

Only a few investigators have used an in vivo model to study the effects of 5-HT on ganglionic transmission. Jones et al. (1995) showed that 5-HT1D receptors mediate inhibition of sympathetic ganglionic transmission in anesthetized cats.

c. Changes in Pathological Condition.

Most recently, 5-HT3 receptors have been suggested to contribute to long-term potentiation in sympathetic ganglia, and inhibition of 5-HT receptors with the antagonist odansetron decreases blood pressure in obese Zucker rats (Gerges et al., 2002). Likewise, tropisetron and odansetron (5-HT3 receptor antagonists) prevented the hypertension caused by psychosocial stress (Alkadhi et al., 2005); the authors suggest that the long-term potentiation facilitated by 5-HT facilitates the increased sympathetic drive that is the basis of the hypertension.

3. Central Effects of 5-Hydroxytryptamine Influencing the Sympathetic Neural Control of Blood Pressure

a. Crossing the blood-brain barrier.

The extent to which 5-HT in the systemic circulation has access to 5-HT receptors on cells within the brain is a significant consideration in understanding the role of 5-HT in regulating blood pressure. In this regard, many of the populations of central neurons controlling the autonomic neural outflows to the heart, vasculature, and kidney have 5-HT receptors and receive inputs from the extensive serotonergic pathways within the CNS or from serotonergic primary afferent neurons (Fig. 5). Thus, depending on the extent to which and the site at which circulating 5-HT crosses the blood-brain barrier, it could influence blood pressure by altering the discharge of 5-HT receptor-expressing neurons in cardiovascular regulatory pathways. In contrast with its precursor, 5-HTP, which moves freely across the blood-brain barrier, the current consensus is that 5-HT does not cross the blood-brain barrier (Hardebo and Owman, 1980; Ohtsuki, 2004; Afergan et al., 2008; Ueno, 2009; Pozdzik et al., 2010). However, a study in 1968 demonstrated that intravenous 5-HT administration results in higher levels of 5-HT and its primary metabolite, 5-HIAA, in the brain (Bulat and Supek, 1968). Moreover, with the finding that SERT is on the capillary endothelial cells that comprise the blood-brain barrier (Brust et al., 2000; Wakayama et al., 2002), and that 5-HT can be transported by these cells (Roux and Couraud, 2005), it is possible for 5-HT to move in and out of the CNS. Nakatani et al. (2008) described the ability of 5-HT to pass from the CNS to the periphery. Westergaard (1975, 1978) demonstrated more than 30 years ago that 5-HT itself can modify blood-brain barrier function, whereas 5-HT may also increase the permeability of the barrier under heat stress (Sharma and Dey, 1984). 5-HT has been localized in the circumventricular organs of the rat (Takeuchi and Sano, 1983). Finally, Alenina et al. (2009) demonstrated that 5-HT was still detectable in the brains of the TPH2(−/−) mouse. This can be interpreted in three ways: 1) that TPH2 knockout is not complete, but there is little other evidence to suggest this is true; 2) there is another source for 5-HT synthesis (such as TPH1 or phenylalanine hydroxylase) that is revealed upon removal of the primary 5-HT synthetic source in the CNS (TPH2); and 3) it is possible that the 5-HT detected in the brain of the TPH2(−/−) mouse is made by TPH1 (located in the periphery), and this 5-HT moves from the periphery to the brain. Bagale et al. (2011) reported the use of ethylamine-functionalized fluorophores that can be used to visualize SERT in living tissues. Thus, it remains possible that systemic 5-HT could have effects on blood pressure by binding to 5-HT receptors on CNS neurons, and we have tools to test this idea. The potential effects of systemic 5-HT acting within the brain can be appreciated from a review of the role of central serotonergic effects on the principal blood pressure-controlling pathways within the brain.

Fig. 5.

Fig. 5.

Central autonomic circuits and potential central sites of action of systemically administered serotonin. The excitatory (+), inhibitory (−), or mixed (±) connections shown between the various autonomic nuclei are based on data described by Barman and Gebber (2000), Guyenet (2006), and Morrison (2004). The solid pentagon symbols show locations of 5-HT receptors. BAT, brown adipose tissue; CVLM, caudal ventrolateral medulla; DMH, dorsomedial hypothalamus; LTF, lateral tegmental field; NA, nucleus ambiguus; NTS, nucleus of the tractus solitarius.

b. Central 5-hydroxytryptamine influences on blood pressure.

The central neural networks regulating blood pressure are primarily those that control the level of sympathetic activities to cardiovascular tissues. Figure 5 shows some of the major neural circuits known to regulate blood pressure and heart rate (Barman and Gebber, 2000; Morrison, 2004; Guyenet, 2006). The sympathetic ganglion cells innervating the heart, blood vessels, adrenal medulla, and kidney receive their principal excitation from sympathetic preganglionic neurons in the intermediolateral nucleus (IML) in the thoracolumbar spinal cord. The sympathetic preganglionic neurons, although influenced by a network of spinal interneurons, receive their primary glutamatergic drive from supraspinal populations of sympathetic premotor neurons, including the major group of cardiovascular regulatory neurons in the rostral ventrolateral medulla (RVLM). The activity of these sympathetic premotor neurons is determined, in turn, by connections from pontine, hypothalamic, and limbic areas and by brainstem inputs from cardiovascular baroreceptor and chemoreceptor reflex circuits. The baroreceptor reflex, for instance, is initiated by sensory input from stretch receptors in the carotid sinus and aortic arch that project to the nucleus of the tractus solitarius (NTS), in a negative feedback fashion, to alter cardiac output and vascular resistance to compensate for externally imposed changes in blood pressure. NTS neurons excite neurons in the caudal ventrolateral medulla that in turn inhibit RVLM sympathoexcitatory neurons. A reduction in baroreceptor activity mediates the sympathetic activation that increases heart rate, cardiac contractility, venous stiffness, and vasoconstriction to compensate for the decrease in blood pressure that occurs upon standing when gravity pulls blood toward the lower extremities. Likewise, baroreceptor sensory neurons in the NTS provide the parasympathetic premotor drive to the cardiac vagal preganglionic neurons in the medulla that drive cardiac vagal nerve activity to slow the heart.

Central serotonergic influences on blood pressure arise from stimulation of serotonin receptors on neurons within these neural circuits that determine sympathetic and vagal outflows (Fig. 5 and Table 3). Not only is there a myriad of central sites at which 5-HT could influence blood pressure, but there are several 5-HT receptor subtypes with differing effects on neuronal activity. Further complicating the interpretation of experimental data is the lack of specificity of many experimental designs in which cardiovascular parameters were measured after administration of 5-HT or related compounds into the cerebral ventricles or into the systemic circulation. Thus, serotonergic drugs can have access to a wide spectrum of 5-HT receptors on diverse populations of neurons. Although this drug delivery approach has relevance to investigations of a compound's therapeutic potential, it is not likely to mimic the normal physiological release of 5-HT from specific subpopulations of 5-HT neurons onto restricted sets of target neurons. It is also apparent, on the basis of the preceding description, that through different receptor populations, 5-HT could have divergent effects on different populations of neurons in the serially connected organization of the central pathways determining the sympathetic and vagal outputs controlling blood pressure. Thus, if 5-HT has simultaneous access to inhibitory 5-HT receptors (such as 5-HT1A receptors) on sympathetic premotor neurons and stimulatory 5-HT receptors (such as 5-HT2A receptors) on sympathetic preganglionic neurons, it is not surprising that administration of serotonin can either inhibit or stimulate preganglionic sympathetic nerves (Lewis and Coote, 1990). Moreover, Pickering et al. (1994) showed that perfusion of a spinal cord slice preparation with 5-HT induced rhythmic activity in sympathetic preganglionic neurons.

TABLE 3.

Effect of acutely administered 5-HT or serotonergic agonist on blood pressure and heart rate in a variety of species

Leftmost column list species or agonist, status of model consciousness, surgical or pharmacological agents on board during experimentation.

5-HT or Agonist and Species Site of Administration Dose Effect on Blood Pressure Effect on Heart Rate Time Point Reference
5-HT
    Rat, conscious IV 2 μg/rat Decrease Decrease Minutes Callera et al., 2005
    Rat, conscious (female) SC 2 mg/kg Decrease Minutes Barney et al., 1981
    Rat, conscious ICV <10 nmol Increase Increase Minutes Dedeoğlu and Fisher, 1991
    Rat, conscious ICV >10 nmol Increase Decrease Minutes Dedeoğlu and Fisher, 1991
    Rat, conscious ICV 4 nmol/kg Increase Decrease Minutes Anderson et al., 1996
    Rat, conscious ICV 10 μg Increase Decrease Minutes Saydoff et al., 1996
    Rat, vagotomized and pithed, ketanserin IV 10−9–10−5 mol/kg Decrease Terrón, 1997
    Rat, vagotomized, anesthetized, ritanserin IV 1–10 μg · kg−1 · min−1 Decrease Minutes De Vries et al., 1999
    Rat, anesthetized, vagosympathecto-mized, ketanserin IV 1–30 μg/kg Decrease Minutes Centurión et al., 2004
    Rat, pithed IV 5–20 μg/kg Increase Minutes Cavero et al., 1981
    Rat, syrosingopine and mianserin IV 20 μg/kg Decrease Minutes Cavero et al., 1981
    Rat, anesthetized IR, IMes 1–3 μg/min Increase Increase Minutes Janssen et al., 1989
    Rat, anesthetized ICV 20 μg Increase Decrease 5 min Montes and Johnson, 1990
    Rat, anesthetized ICV 20 μg Decrease Decrease 10 min Montes and Johnson, 1990
    Rat, anesthetized ICV 40, 120 nmol/kg Increase Decrease Minutes Anderson et al., 1992
    Rat, anesthetized RVLM 5–50 nmol Decrease Decrease Minutes Key and Wigfield, 1992
    Cat, anesthetized Fourth ventricle 20–640 nmol/kg Decrease Decrease Minutes Shepheard et al., 1994
    Dog, anesthetized IV 8 μg · kg−1 · min−1 Decrease Increase Minutes-hours Carlson et al., 1967.
    Dog, anesthetized IV 20 μg · kg−1 · min−1 Decrease Increase Minutes Martinez and Lokhandwala, 1980
    Calf, metrenperone IV 0.05 mg · kg−1 · min−1 Decrease Decrease, then increase Minutes Linden et al., 1999
    Human, conscious IV 2–4 μg · kg−1 · min−1 Variable Minutes–hours Carlson et al., 1967
    Human, conscious IV 20 nmol · kg−1 · min−1 No Change No Change Minutes Hansen et al., 2008
    Human, conscious IL 2.5–250 nmol No Change No Change Minutes Hansen et al., 2008
    Human, conscious IA (forearm) 10–80 ng · kg−1 · min−1 No change No change Minutes Blauw et al., 1988
8-OH-DPAT
    Rat, anesthetized IV 8–128 μg/kg Decrease Decrease Minutes-hours Fozard et al., 1987
    Rat, anesthetized IV 1 mg/kg Decrease Decrease Minutes Helke et al., 1993
    Rat, conscious IP 0.05–0.25 mg/kg Decrease Minutes van den Buuse and Wegener, 2005
    Rat, conscious, hemorrhaged ICV, IV 48 nmol/kg Increase Minutes Scrogin, 2003
    Rat, conscious, hemorrhaged IV 30 nmol/kg Increase Minutes Tiniakov et al., 2007
    Rat, conscious ICV <10 nmol Increase Increase Minutes Dedeoğlu and Fisher, 1991
    Rat, conscious ICV >10 nmol Decrease Decrease Minutes Dedeoğlu and Fisher, 1991
    Rat, hemorrhaged ICV 10 nmol Increase Increase Minutes Jochem et al., 2009
    Rat, anesthetized Ventral medulla 0.1 μg/0.1 μl Decrease Decrease Minutes Valenta and Singer, 1990
    Rat, anesthetized ICV 3 nmol/kg Increase Increase Minutes Anderson et al., 1992
    Rat, anesthetized Ventral medulla 20–50 ng Decrease Decrease Minutes Helke et al., 1993
    Rabbit, hypovolemic conscious Fourth ventricle 10–30 nmol Increase Minutes Evans et al., 1993
    Cat, anesthetized IV 1–100 μg/kg Decrease Decrease Minutes Ramage and Fozard, 1987
    Cat, anesthetized Fourth ventricle 2.5–40 nmol/kg Decrease Decrease Minutes Shepheard et al, 1994
    Pig IA 0.3–10 μg · kg−1 · min−1 None Decrease Minutes Bom et al., 1989
    Dog IV 10–300 μg/kg Decrease No change Minutes Dabiré et al., 1990
    Dog VA 0–3 μg/kg Decrease Decrease Minutes Dabiré et al., 1990
    Dog VLPA 0.2 μg/site Decrease Decrease Dabiré et al., 1990
Buspirone
    Rat, conscious IP 0.1/0.5 mg/kg Decrease/increase Minutes van den Buuse and Wegener, 2005
5-CT
    Rat, anesthetized, vagotomized. ketanserin IV 0.01–0.3 μg/kg Decrease Minutes Centurión et al, 2004
    Rat, vagotomized, pithed, ketanserin IV 10−11-10−7 mol/kg Decrease Minutes Terrón, 1997
    Rat, anesthetized ICV 3 nmol/kg Increase Increase Minutes Anderson et al., 1992
    Cat, anesthetized Fourth ventricle 2.5–40 nmol/kg Decrease No change Minutes Shepheard et al, 1994
DP-5-CT
    Rat, anesthetized ICV 3 nmol/kg Increase Increase Minutes Anderson et al., 1992
    Cat, anesthetized Fourth ventricle 2.5–40 nmol/kg Decrease Decrease Minutes Shepheard et al., 1994
Rizatriptan
    Rat, anesthetized IV 0.63–2500 μg/kg Decrease Decrease Minutes Pagniez et al., 1998
Sumatriptan
    Rat, anesthetized IV 0.63–2500 μg/kg Decrease Decrease Minutes Pagniez et al., 1998
    Cat, anesthetized Fourth ventricle 10–160 nmol/kg Decrease No change Minutes Shepheard et al., 1994
α-Methyl-5-HT
    Rat IV 3–30 μg Increase Minutes Dalton et al., 1986
5-Methoxytryptamine
    Rat, vagotomized, pithed, ketanserin IV 10−9–10−5 mol/kg Decrease Minutes Terrón, 1997
DOI
    Rat, anesthetized ICV 40, 120 nmol/kg Increase Decrease Minutes Anderson et al., 1992
    Rat, anesthetized Intra-NTS 0.05–1 pmol Decrease Decrease Minutes Comet et al., 2007
    Rat, anesthetized ICV 2 μmol/kg Increase Increase Minutes Knowles and Ramage, 1999
DOB
    Rat, anesthetized Intra-NTS 0.025–0.5 pmol Decrease Decrease Minutes Merahi and Laguzzi, 1995
m-CPBG
    Rat, anesthetized Third ventricle 80–320 nmol Decrease No change Minutes Ferreira et al., 2004
2-Methyl-5-HT
    Rat, anesthetized IV 3–30 μg Decrease Decrease Minutes Dalton et al., 1986
m-CPP
    Rat, anesthetized Third ventricle 80–320 nmol Increase Decrease then increase ≤30 min Ferreira et al., 2005
Quipazine
    Rat, anesthetized ICV 2 μmol/kg Increase Increase Minutes Knowles and Ramage, 1999
5-HTP
    Rat, conscious IV 2–40 mg · kg−1 · h−1 Decrease 24 h Echizen and Freed, 1982
    Cat, anesthetized IV 5–10 mg/kg Decrease Hours Flórez and Armijo, 1974
Tryptophan
    Rat, SHR, conscious IP 1–100 mg/kg Decrease Hours Sved et al., 1982
    Rat, conscious IV 2–40 mg · kg−1 · h−1 No change 24 h Echizen and Freed, 1982

5-HTP, 5-hydroxytryptophan; DOB, 2,5-dimethoxy-4-bromoamphetamine hydrochloride; DOI, (±)-2,5-dimethoxy-4-iodoamphetamine; DP-5-CT, dipropyl-5-carboxamidotryptamine; IA, intraarterial; ICV, intracerebroventricular; IL, intraluminal (intestinal); IMes, intramesenteric artery; IP, intraperitoneal; IR, intrarenal artery; IV, intravenous; m-CPBG, m-chlorophenylbiguanidine; m-CPP, 1-(3-chlorophenyl)piperazine hydrochloride; VA, vertebral artery; VLPA, ventrolateral pressor area.

Central stimulation of 5-HT1A-receptors reduces vasoconstrictor sympathetic nerve activity and increases cardiac vagal nerve activity, both leading to a decrease in blood pressure (Ramage, 2001). This may occur through inhibition of vasomotor premotor neurons in the RVLM (McCall and Clement, 1994), although a potential role of adrenergic receptors in this effect has been suggested (Nosjean and Guyenet, 1991), or of their antecedent sympathoexcitatory neurons in the lateral tegmental field (McCall and Clement, 1994). As shown in Fig. 5, sympathetic premotor neurons controlling cutaneous vasoconstriction are located in the rostral medullary raphe, where a local injection of a 5-HT1A-receptor agonist elicits a cutaneous vasodilation (Ootsuka and Blessing, 2006). Because the cutaneous vasculature is strongly constricted at normal ambient temperatures, inhibition of the skin sympathetic outflow could contribute significantly to a 5-HT1A-receptor agonist-evoked decrease in blood pressure. 5-HT1A-receptor agonists in the spinal IML can, however, elicit a sympathoexcitation as a result of increased responsiveness to glutamatergic inputs (Madden and Morrison, 2006), potentially through inhibition of local GABA neurons.

Central stimulation of 5-HT2A-receptors can lead to an increase in blood pressure, partly through increased vasoconstrictor sympathetic outflow arising from activation of sympathetic premotor neurons in the RVLM (Ramage and Daly, 1998) but also from the contribution of vasopressin release (Saydoff et al., 1996). Intravenous administration of a 5-HT2A-receptor agonist facilitates the cutaneous sympathetic outflow (Blessing and Seaman, 2003), potentially at the IML site of sympathetic preganglionic neurons (Ootsuka and Blessing, 2005), and thus could contribute to an elevated arterial pressure in response to central 5-HT2A-receptor agonist administration.

The NTS is the site of termination of primary sensory neurons, including those from the baroreceptors, chemoreceptors, and cardiopulmonary receptors that participate in reflex regulation of blood pressure. The NTS receives serotonergic inputs from serotonergic neurons in the nodose ganglia and from neurons in the medullary raphe nuclei (Thor and Helke, 1987; Nosjean et al., 1990) (Fig. 5), and 5-HT1A, 5-HT1B, 5-HT2, 5-HT3, and 5-HT4 receptors have been identified within the NTS, including on vagal afferent nerve terminals (Laguzzi, 2003).

Microinjection of 5-HT into the NTS can elicit either depressor or pressor responses (Feldman and Galiano, 1995; Nosjean et al., 1995; Callera et al., 1997; Laguzzi, 2003) that could arise from facilitatory or inhibitory effects of 5-HT on the NTS circuitry involved in integrating the baroreceptor and other cardiovascular reflexes. Chemical destruction of serotonergic neuronal elements in the NTS by microinjection of the neurotoxin 5,7-dihydroxytryptamine led to a transient (over 6 days) increase in blood pressure (Orer et al., 1991). Selective activation of 5-HT2A receptors in the NTS by microinjection of DOI significantly enhanced the cardiovagal component of the baroreceptor reflex (Laguzzi, 2003), consistent with the possibility that, under physiological conditions, 5-HT released from the projections originating in the nodose ganglia and/or nucleus raphe pallidus might trigger the 5-HT2A receptor-mediated reflex responses.

On the other hand, microinjection of 5-HT or the selective 5-HT3 receptor agonist 1-(m-chlorophenyl)-biguanide into the rat NTS increases lumbar sympathetic nerve activity and blood pressure (Nosjean et al.,1995), which was prevented by prior microinjection of zacopride, a 5-HT3 receptor antagonist. Because the gain of the sympathetic component of the baroreceptor reflex was not changed, they concluded that activation of 5-HT3 receptors in the NTS causes sympathoexcitation by a mechanism independent of the baroreceptor reflex. In this regard, 5-HT3 receptor activation excites neurons in the NTS and in the dorsal motor nucleus of the vagus, directly adjoining the NTS, by a glutamate-dependent mechanism (Jordan, 2005), and 5-HT3 receptor blockade reduces excitatory amino acid transmission in NTS (Wan and Browning, 2008). Vagal afferents expressing 5-HT3 receptors arise not only from the gut but also from the heart, mediating the Bezold-Jarish reflex. Activation of 5-HT3 receptors on vagal afferents could influence blood pressure through a reduction in the effectiveness of the arterial chemoreceptor reflex (Moreira et al., 2007). Thus, the effects on blood pressure of 5-HT administration into the NTS and its effects on specific reflexes that influence blood pressure is complex, varying with factors such as anesthesia, species, and relative prominence of or exposure to various 5-HT receptor subtypes. The existence of 5-HT receptors on both the sensory and synaptic terminals of primary afferents suggests that transmission in these pathways could be influenced both by the levels of circulating 5-HT and by 5-HT released within the NTS from the terminals of raphe neurons.

V. Effect of 5-Hydroxytryptamine on Blood Pressure: Whole-Animal Studies

A. Effects of 5-Hydroxytryptamine and Serotonergic Agonists on Blood Pressure

1. Short-Term.

When 5-HT is administered intravenously to anesthetized rodents over the course of seconds to minutes, a classic triphasic response is observed (Dalton et al., 1986; Hardcastle and Hardcastle, 1999) (Fig. 6A). There is 1) a depressor response attributed to a reduction in heart rate (HR) by activation of the Bezold-Jarisch reflex followed by 2) a significant elevation in blood pressure. This pressor response is thought to be mediated by 5-HT2 receptors in the vasculature. After this transient pressor response, there is 3) a slow vasodepressor response that has been attributed to activation of both 5-HT7 and 5-HT1B/1D receptors (Terrón et al., 2007). In studies of anesthetized animals, this response can be observed up to 60 min after 5-HT administration and a decrease in blood pressure to 5-HT can continue for at least a week with continued 5-HT administration.

Fig. 6.

Fig. 6.

A, classic triphasic (1, 2, 3) effect of 5-HT (75 μg/kg bolus) on blood pressure in the anesthetized rat. Time base below traces is 1 s/division. Total time base is 75 s. B, effect of 5-HT (25 μg · kg−1 · min−1, subcutaneous pump) on blood pressure (top) and heart rate (bottom) in the conscious rat over the course of 30 h. Dashed vertical line indicates implantation of pump. Points represent means ± S.E.M. for the number of animals in parentheses.

Table 3 compiles responses to 5-HT administered in a number of different ways, but all in an acute (seconds to minute) time frame. Multiple investigators have demonstrated the ability of 5-HT, given either subcutaneously or intravenously, to decrease blood pressure in either rodents or dogs. In a pithed rat model, administration of 5-HT results in an increase in blood pressure (Cavero et al., 1981), suggesting that the interaction with the central sympathetic nervous system may be key for the mechanism of blood pressure decrease. There are few reports of the response of the human to 5-HT (intravenous), and in one study that addressed directly the effect of 5-HT on blood pressure in man, the effects of 5-HT on blood pressure were variable (Carlson et al., 1967). 5-HT (0.1–50 ng · kg−1 · min−1, intra-arterial) caused a dose-dependent increase in forearm blood flow in humans, but this alone was insufficient to decrease blood pressure, which is somewhat expected given the restriction of circulating substances in the assay used (Blauw et al., 1988).

The effect of serotonergic agonists other than 5-HT on blood pressure are tabulated in the lower part of Table 3. In the conscious rat, the 5-HT1A receptor agonist 8-OH-DPAT causes variable effects on blood pressure when given intraperitoneally or intravenously but overall reduces blood pressure (van den Buuse and Wegener, 2005). By contrast, in the hemorrhaged rat, 8-OH-DPAT produces a significant pressor response and is described as rescuing blood pressure from the effects of hemorrhage (Tiniakov et al., 2007). In the cat (anesthetized), 8-OH-DPAT (intravenous) causes a decrease in blood pressure (Ramage and Fozard, 1987). 5-CT has a significantly greater affinity for more of the 5-HT1 receptor subtypes and causes a decrease in blood pressure (Terrón et al., 2007), with the caveat that 5-CT also has significant affinity for the 5-HT7 receptor (Waeber and Moskowitz, 1995; Krobert et al., 2001). These animals, however, were anesthetized, vagotomized, and treated with ketanserin to unmask a 5-HT-stimulated relaxation/depressor response (Terrón, 1997; Centurión et al., 2004), so, as with isolated arteries, the physiological relevance of such a response remains in question. Sumatriptan and rizatriptan (intravenous), 5-HT1B/1D receptor agonists, also caused a decrease in blood pressure, but this effect may not be peripheral. Pagniez et al. (1998) have suggested that this effect is due to central 5-HT1A receptor activation. In contrast to these studies, in which serotonergic agonists were clearly depressors when given peripherally, 5-HT2 receptor agonists such as α-methyl-5-HT and (±)-2,5-dimethoxy-4-iodoamphetamine elevate blood pressure. Because it is unclear whether these agonists—including 5-HT—cross the blood-brain barrier or interact with the area postrema, we cannot exclude the idea that the effects observed on blood pressure are at least partially centrally mediated.

2. Long-Term.

Because of our interest in the role of 5-HT in blood pressure, we studied prolonged elevation of 5-HT in the normal rat and followed blood pressure using radiotelemetry. The time period of 5-HT infusion was over the course of a week. Although many of the studies described in Table 3 suggest that 5-HT is largely a depressor agent in short-term peripheral administration, the knowledge that 1) 5-HT is a direct vasoconstrictor, 2) blood vessels from hypertensive animals are hyper-responsive to 5-HT, and 3) circulating levels of free 5-HT are elevated in hypertension suggested to us that 5-HT could be a pathogenic factor in hypertension, contributing to an elevation in blood pressure. Contrary to these ideas, we observed that 5-HT caused a decrease in blood pressure in the conscious sham and DOCA-salt hypertensive rats (Diaz et al., 2008). In these studies, 5-HT was given in an Alzet miniosmotic pump (25 μg · kg−1 · min−1) placed subcutaneously and with an antioxidant. On day 7, plasma 5-HT (free) was elevated from 2.7 ± 0.03 (vehicle) to 47.1 ± 13.18 ng/ml (5-HT; 17-fold increase) in a sham rat. This is within the range of that found in experimental and genetic hypertensive rats (25–60 ng/ml or ∼125 nM 5-HT). Measures of human 5-HT have been carefully re-evaluated using liquid chromatography–tandem mass spectrometry. Estimated 5-HT concentrations in platelet-depleted plasma were 89.5 to 115.5 nM in 18 healthy subjects (Monaghan et al., 2009). Thus, the rat is a good model to use because the levels of 5-HT we achieve in this model are physiological and comparable with what is found in human. Stunningly, 5-HT nearly normalized the blood pressure of the DOCA-salt hypertensive rat (Table 4), dropping blood pressure by over 50 mm Hg and maintaining a lower pressure for the entire week. Platelet free 5-HT levels were increased 10- to 15-fold in the experiments represented in these tables. This was the first study to demonstrate that, unlike the effects of 5-HT in minutes of infusion, 5-HT continues to exert its antihypertensive effects over days. It is possible but not proven that with the longer-term infusion, we are extending phase 3 of the response to 5-HT (Fig. 6A, phase 3). To investigate this possibility, we tracked changes in blood pressure and heart rate within the first 30 h after 5-HT administration in normal male Sprague-Dawley rats (Fig. 6B, conscious). Our findings suggest that there are at least two hypotensive phases upon long-term 5-HT administration: one within the first 3 h and a second, more stable decrease after nearly 20 h of administration. Heart rate remains elevated during these first 30 h. Thus, it is unlikely that the long-term (1 week) decrease in blood pressure to 5-HT is simply an extension of the decrease observed within the first hour. We do not know whether these different phases of changes in blood pressure have the same mechanism. In Fig. 7, we demonstrate the ability of 5-HT [Alzet miniosmotic pump (25 μg · kg−1 · min−1)] to lower blood pressure in the spontaneously hypertensive rat (SHR) over 1 week. Blood pressure was significantly reduced within the first 2 days of delivery, during which time the HR was elevated. HR returned to normal levels, whereas blood pressure remained reduced by day 7. This suggests that although the baroreceptor reflex is intact, it cannot completely correct for the decrease in pressure. Table 4 compiles the results of experiments from our laboratory investigating the response to a 1-week infusion of 5-HT [Alzet miniosmotic pump (25 μg · kg−1 · min−1)] on blood pressure in several rodent models, and these studies support the ability of 5-HT to reduce rodent blood pressure.

TABLE 4.

Effect of long-term administration of 5-HT (25 μg · kg−1 · min−1, 1-week administration, subcutaneous) on blood pressure in the rat

Rat Group (All n > 4) Control BP Nadir BP Time Point Reference
mm Hg h
Male SD (CRiver) 101 ± 2 86 ± 1 24 Diaz et al., 2008
Male SD (CRiver) 102 ± 3 79 ± 2 48 Diaz et al., 2008
Male DOCA-salt SD (CRiver) 166 ± 7 112 ± 3 48 Diaz et al., 2008
Male SD (Harlan) 103 ± 3 83 ± 3 24 Diaz et al., 2008
Male l-NNA SD (Harlan) 161 ± 4 153 ± 5 72 Diaz et al., 2008
Male SERT(+/+) (Wistar) 107 ± 11 84 ± 1 24 Davis et al., 2011
Male SERT(−/−) (Wistar) 103 ± 2 88 ± 4 24 Davis et al., 2011
Female SERT(+/+) (Wistar) 111 ± 3 89 ± 2 24 Davis et al., 2011
Female SERT(−/−) (Wistar) 109 ± 2 94 ± 3 24 Davis et al., 2011

BP, blood pressure; CRiver, Charles River Laboratories; SD, Sprague Dawley.

Fig. 7.

Fig. 7.

Top, ability of 5-HT (25 μg · kg−1 · min−1, subcutaneous pump) to lower blood pressure of a male SHR. Bottom, concomitant heart rate measures. Dashed vertical line indicates implantation of the pump. Points are means ± S.E.M. for the number of animals in parentheses. *, p < 0.05, significantly different from vehicle time point.

Potential mechanisms contributing to the 5-HT-induced decrease in blood pressure in the chronic setting include a role for NOS and for the SERT. The NOS inhibitor l-NNA abolished the ability of 5-HT to cause a decrease in blood pressure in both normal and DOCA-salt rats given l-NNA. We have validated this finding in a study using newly available iPrecio pumps (Primetech Corp., Tokyo, Japan), which allow for programmable release of infusions over time. We conducted a dose-response curve to 5-HT within normal rats and demonstrated a dose-dependent decrease in blood pressure, one that could be prevented with administration of l-NNA (Tan et al., 2011). Thus, 5-HT is interacting with NOS somewhere to effect the decrease in blood pressure. The decrease in 5-HT is not physiologically antagonized by l-NNA, which elevated the initial blood pressure, because 5-HT readily reduces the blood pressure of the DOCA-salt and SHRs (Diaz et al., 2008; Fig. 4). 5-HT can increase NO production from several different types of cells and tissues (Arima et al., 1996; Manivet et al., 2000; Borgdorff et al., 2002; Bellou et al., 2003; Borgdorff and Tangelder, 2006; García et al., 2006; Chanrion et al., 2007). Hoffmann et al. (1990) described 5-HT receptors as mediating the long-lasting decrease in blood pressure after exercise and implicated 5-HT1 and 5-HT2 receptors in mediating this response. Use of the SERT dysfunctional rat also suggests that SERT function is important to 5-HT-induced decrease in blood pressure because the magnitude of the blood pressure decrease induced by 5-HT was reduced by 50% in the SERT dysfunctional mice compared with the male wild-type mice (Davis et al., 2011). SERT function is important in the concentration of 5-HT by blood vessels (Ni et al., 2004), but the presence of SERT expression on the blood-brain barrier suggests the possibility that 5-HT crosses the blood-brain barrier to cause a decrease in blood pressure.

This work with long-term 5-HT administration comes approximately 30 years after a series of elegant studies done using 5-HTP and investigating the effect of this 5-HT precursor on blood pressure. 5-HTP but not tryptophan lowered blood pressure in a prolonged (24–48 h) fashion (Echizen and Freed, 1982; Fregly et al., 1987a; Ding et al., 1989; Itskovitz et al., 1989; Baron et al., 1991) (Table 3). 5-HTP is dedicated to 5-HT synthesis, whereas only 5 to 10% of tryptophan is dedicated to 5-HT. A study by Sved et al. (1982) differs in that tryptophan (intraperitoneal) was able to reduce blood pressure of SHRs over the course of hours. Likewise, Fregly et al. (1987b) demonstrated that tryptophan administration inhibited the development of DOCA-salt hypertension in the rat. 5-HTP also inhibited the development of DOCA-salt hypertension (Fregly et al., 1987a) and thus raises the possibility that 5-HT is able to reduce the activity of mechanisms responsible for the elevation of blood pressure. Like tryptophan, 5-HTP is relevant because of the enormous concern of the side effects of unregulated administration of 5-HTP for aide in a multitude of disorders, including depression, fibromyalgia, obesity, insomnia, and headache (Birdsall, 1998; Das et al., 2004; Turner et al., 2006).

C. Effect of Removal of 5-Hydroxytryptamine on Blood Pressure

Another way to ask the question of the importance of 5-HT to blood pressure is to investigate the outcome of removing 5-HT from the body or preventing its synthesis. This has been done in two ways: pharmacological inhibition of TPH by use of parachlorophenylalanine (PCPA) and removal of the TPH gene. p-Chlorophenylalanine methyl ester (PCPAME) was used for decades as a tool to inhibit TPH. In 1976, Buckingham et al. demonstrated that PCPAME (400 mg/kg) caused a significant decrease in blood pressure that lasted for at least 8 days in the male DOCA-salt hypertensive rat, with the reduction in blood pressure observed on the first day of administration. Although this decrease was present in the normal animal, it did not reach the magnitude of decrease reached in the DOCA-rat and was observed only after 5 days of treatment (a latent response). Brainstem 5-HT was depleted by PCPAME, but circulating 5-HT was not measured. These findings suggest that 5-HT promotes a rise in blood pressure by an unknown mechanism, and this is not consistent with the findings just presented in which long-term 5-HT exposure caused a reduction in blood pressure. By contrast, other studies suggest that depletion of 5-HT centrally (primarily brainstem) results in an elevation in blood pressure. Central PCPA administration to Wistar rats was accompanied by elevated blood pressure under anesthesia (Althaus et al., 1985), and central depletion of 5-HT was again associated with an elevated blood pressure in awake rats (Kellett et al., 2005).

Similarly conflicting outcomes have been observed when 5-HT synthesis is removed genetically. TPH1 knockout mice have been created; this isoform of TPH is primarily responsible for peripheral 5-HT, and these mice have normal central 5-HT levels. Systemic arterial pressure (carotid catheterization) was measured in halothane-anesthetized TPH1 KO mice and, compared with WT, TPH1 KO mice showed a significant elevation in basal arterial pressure, where pressure was taken in anesthetized mice (Morecroft et al., 2007). This implies that peripheral 5-HT acts to lower blood pressure or inhibits a mechanism that supports blood pressure. It would have been ideal to measure blood pressure using telemetry, given that anesthetics themselves typically lower blood pressure. Significantly more work has been done with the TPH2 KO mouse, a mouse in which removal of 5-HT is primarily in the central nervous system, and peripheral 5-HT is largely normal. Blood pressure measured telemetrically in TPH2 KO mice was significantly lower than that of the wild type, especially during late afternoon and evening (Alenina et al., 2009). A caveat of this study is that activity was not measured, and because decreased activity (as may be seen with extended sleep) is associated with a decreased blood pressure, the blood pressure data are somewhat difficult to interpret. Brain 5-HT was not abolished in the TPH2 KO mice but significantly reduced. Taken together, 5-HT seems to have different roles in regulating blood pressure by acting in the periphery (lower) versus central compartment (elevation), at least in the mouse. Mice in which both TPH1 and TPH2 are knocked out have been created (Savelieva et al., 2008), but blood pressure in these animals was not measured. As expected, TPH1 KO but not TPH2 KO mice show dramatically reduced levels of circulating 5-HT. It is noteworthy that measures of peripheral levels of 5-HT in the blood were not zero in mice in which both TPH1 and TPH2 were knocked out. The fact that 5-HT levels were not zero in these animals is interesting, and it has been proposed that enzymes such as phenylalanine hydroxylase may serve to produce 5-HT (Savelieva et al., 2008).

VI. 5-Hydroxytryptamine, Pharmaceutical Compounds and Clinical Situations

A. Pharmaceuticals

Drugs that modify 5-HT concentration are frequently prescribed. These drugs prevent either the uptake or the metabolism of 5-HT, both of which would promote the elevation of 5-HT concentration. As such, study of the effects of these drugs on blood pressure may give additional insights into the potential actions of 5-HT in modifying blood pressure. To our knowledge, no drugs that inhibit 5-HT production are currently used therapeutically. A selective serotonin-reuptake enhancer has been described—tianeptine—but it is the only one of its class and its effectiveness as an enhancer of SERT has been questioned. Lexicon Pharmaceuticals (The Woodlands, TX) currently has the tryptophan hydroxylase inhibitor telotristat etiprate (LX1032) in phase 2 clinical trials for carcinoid tumors, a condition in which 5-HT is made in inappropriately high amounts by slow-growing tumors of the enterochromaffin cells in the intestine. It is noteworthy that hypotension is one of the presenting symptoms of patients with this type of tumor.

The various classes of drugs that modify 5-HT concentration include the selective serotonin-reuptake inhibitors (SSRI), serotonin/norepinephrine-reuptake inhibitors (SNRI), and MAO A/B monoamine oxidase. The oldest class of drugs is the tricyclic antidepressants (TCAs). A preponderance of these drugs are/were used in the treatment of depression and some related mood disorders. Table 5 compiles data from a number of studies, animal and human, in which mean arterial blood pressure was measured as one endpoint. Blood pressure was either the focus of the study cited or a variable measured within a trial. Relatively clear patterns emerge from the gathering of these studies. Use of TCAs—still prescribed—has been associated directly with a decrease in blood pressure and/or a decreased tolerance for blood pressure changes associated with moving from a supine to standing posture (orthostatic hypotension). This finding was consistent in multiple species tested, including dog, rabbit, and human. MAO inhibitors have also been used clinically; they, too, are associated with a decrease in blood pressure or orthostatic hypotension (Stahl and Felker, 2008). Some of the later compounds, such as rasagiline, seem to exert less of an effect on blood pressure.

TABLE 5.

Effect of serotonergic compounds on blood pressure of different species

Compound Species Effect on Blood Pressure Dose Reference
TCAs
    Imipramine Human Orthostatic hypotension 1–5 g/kg/day Giardina et al., 1985
    Imipramine Human Orthostatic hypotension 100 mg/day Thayssen et al., 1981
    Imipramine Human Orthostatic hypotension ∼225 mg/day Glassman et al., 1979
    Imipramine Rabbit Decrease 29 mg/kg Hughes and Radwan, 1979
    Imipramine Dog Decrease 0.1–10 mg/kg Kato et al., 1974
    Imipramine Dog Decrease 1–10 mg/kg Yokota et al., 1987
    Imipramine Human Orthostatic hypotension 200 mg/day Guelfi et al., 1983
    Imipramine Dog Decrease 0.5 mg · kg−1 · min−1 Lindbom et al., 1982
    Nortriptyline Human Orthostatic hypotension 0.5–3.5 mg/kg/day Giardina et al., 1985
    Nortriptyline Human No change 40 mg/day Thayssen et al., 1981
    Amitriptyline Rabbit Decrease 15 mg/kg Hughes and Radwan, 1979
    Amitriptyline Dog Decrease 0–20 mg/kg Lindbom et al., 1982
    Amitriptyline Dog Decrease 1–10 mg/kg Yokota et al., 1987
    Amitriptyline Dog Decrease 0.1–10 mg/kg Kato et al., 1974
    Amitriptyline Human Decrease 25 mg/day Ogura et al., 1983
    Amitriptyline Human No change 75 mg/day Penttilä et al., 2001
    Maprotiline Rabbit Decrease 34 mg/kg Hughes and Radwan, 1979
    Clomipramine Dog Decrease 1–10 mg/kg Yokota et al., 1987
    Clomipramine Human Orthosatic hypotension 150 mg/day Christensen et al., 1985
    Clomipramine Dog Decrease 0–20 mg/kg Lindbom et al., 1982
    Desmethylimipramine Human Increase 100 mg Ross et al,. 1983
    Desipramine Human No change 100 mg/day Chalon et al., 2003
MAO Inhibitors
    Rasagiline Rat No change 0.2–1 mg/kg Finberg et al., 2006
    Rasagiline Rat No change 1 mg/kg Abassi et al., 2004
    Rasagiline Rat Decrease 10 mg/kg Abassi et al., 2004
    Selegiline Rat No change 1–5 mg/kg Finberg et al., 2006
    Selegiline Human Orthostatic hypotension 10 mg/day Turkka et al., 1997
    Selegiline Rat No change 1 mg/kg Abassi et al., 2004
    Selegiline Rat Decrease 10 mg/kg Abassi et al., 2004
    L-Deprenyl Rat No change 5 mg/kg Stevens et al., 1998
    L-Deprenyl Rat Increase 25 μg/100 g Kerecsen and Bunag, 1989
    Clorgyline Rat No change 2 mg/kg Finberg et al., 2006
    Clorgyline Rat No change/decrease 25 μg/100g Kerecsen and Bunag, 1989
    Clorgyline Human Decrease 28 mg/day Murphy et al., 1979
    Harmaline Rat Decrease 20 mg/kg Marwood et al., 1985
    Pargyline Rat Decrease 100 mg/kg Marwood et al., 1985
    Pargyline Rat Decrease 10 mg/kg Fuentes et al., 1979
    Pargyline Human Decrease 87 mg/day Murphy et al., 1979
    Tranylcypromine Rat Decrease 10 mg/kg Marwood et al., 1985
    Tranylcypromine Rat Decrease 5 mg/kg Ashkenazi et al., 1983
    Tranylcypromine Human Orthostatic hypotension 20–100 mg/day Nolen et al., 1993
    Tranylcypromine Rat Increase 25 mg/kg Finberg et al., 2006
    Isoniazid Rat Decrease 30–300 mg/kg Vidrio et al., 2000
    Brofaromine Human Orthostatic hypotension 50–250 mg/day Nolen et al., 1993
SSRIs
    Citalopram Human No change 40 mg/day Christensen et al., 1985
    Citalopram Human No change 40–60 mg/day Pedersen et al., 1982
    Citalopram Human No change 20 mg/day Penttilä et al., 2001
    Fluvoxamine Human No change 300 mg/day Guelfi et al., 1983
    Fluoxetine Rat Increase 10–50 μg i.c.v. Lazartigues et al., 2000
    Fluoxetine Rat Decrease 10 mg/kg Sved et al., 1982
    Fluoxetine Human Increase 20 mg/day Amsterdam et al., 1999
    Paroxetine Dog Increase/ Decrease 0.1–10 mg/kg Yokota et al., 1987
    Sertraline Human, adolescent No change Up to 200 mg/day Wilens et al., 1999
    Sertraline Human No change Up to 400 mg/day Saletu et al., 1986
    Zimeldine Dog Decrease 1–20 mg/kg Lindbom et al., 1982
    Zimeldine Human No change 100 mg/day Saletu et al., 1986
SNRIs
    Milnacipran Human Increase 0.2–0.8 mg/kg Caron et al., 1993
    Venlafaxine Human Increase 75–225 mg/day Perahia et al., 2008
    Venlafaxine Human Decrease 75–225 mg/day Alexandrino-Silva et al., 2008
    Venlafaxine Human Increase 225–525 mg/day Mbaya et al., 2007
    Duloxetine Rat Decrease 30 mg/day Chudasama and Bhatt, 2009
    Duloxetine Human Increase 120–400 mg/day Derby et al., 2007
    Duloxetine Human No change 60–120 mg/day Perahia et al., 2008
    Duloxetine Human No change 40–120 mg/day Bailey et al., 2006
    Duloxetine Human Increase 80 mg/day Chalon et al., 2003
    Sibutramine Human Decrease 10–15 mg/day Sharma et al., 2009
    Sibutramine Human Basal increase, decrease in sympathetic stimulation 15 mg/day Birkenfeld et al., 2005
    Sibutramine Rat No change 5 mg/day Chudasama and Bhatt, 2009
    Sibutramine Human Decrease 10–15 mg/day Gaciong and Placha, 2005
    Sibutramine Human Increase 5–20 mg/day McMahon et al., 2000

With the introduction of zimeldine, indalpine, fluvoxamine, and fluoxetine in the mid- to late 1980s, the class of SSRIs was born. Zimeldine was the first developed SSRI (Carlsson and Wong, 1997), and fluoxetine (Prozac; Eli Lilly & Co., Indianapolis, IN) is the best-known SSRI. There is little information as to the effect of fluoxetine, or a related compound fluvoxamine, on blood pressure except for reports in the rat, in which fluoxetine causes hypertension (Tsai and Lin, 1986; Lazartigues et al., 2000). By contrast, fluoxetine reduced blood pressure in the SHR (Sved et al., 1982). It is noteworthy that fluoxetine has been cited for use in the treatment of orthostatic hypotension, the very syndrome caused by TCAs and MAO inhibitors (Grubb et al., 1994). Thus, fluoxetine has divergent effects on blood pressure. Other compounds in this class—citalopram, paroxetine, sertraline, and zimeldine—have varied effects on blood pressure, no effect being the most common outcome. This differs significantly from the effect of SNRIs. Table 5 lists some of the reports that suggest this class of drugs has a more pronounced effect on blood pressure, elevating blood pressure in many cases. A number of meta-analyses have been performed in this regard (Thase, 1998; Kim et al., 2003; de Lemos et al., 2008; Johansson et al., 2010). Because these compounds inhibit reuptake of NE as well as 5-HT, one can understand their potential to elevate blood pressure, because the SNRI would prolong the effect of NE (and 5-HT) at the vascular junction to promote an increase in total peripheral resistance. The SNRI sibutramine was a drug prescribed for medical management of weight loss but was pulled from the market in 2010 because of concerns over cardiovascular side effects that include an elevation in blood pressure. It is less clear how SNRIs would reduce blood pressure, but such a result has been observed in humans for both venlafaxine and duloxetine.

On the whole, it is difficult to attribute the effects of any of the MAOIs, SSRIs, SNRIs, or TCAs to inhibition of metabolism or uptake of 5-HT because of the importance of MAO and uptake systems to norepinephrine and epinephrine. MAO metabolizes both amines as well as 5-HT, and the transport systems are known to be imperfect in their selectivity, such that NET can take up 5-HT and vice versa (Daws, 2009). Moreover, whereas these compounds each elevate 5-HT concentration, their effects are not absolutely specific for 5-HT. For example, more than a dozen publications describe the ability of fluoxetine and its metabolite norfluoxetine to inhibit multiple classes of potassium channels that operate in the cardiovascular system (exemplified by Tytgat et al., 1997). Given our experimental findings of the ability of 5-HT to reduce blood pressure in the long term, it is tempting to speculate that the decrease in blood pressure observed in response to these drugs is due specifically to inhibition of 5-HT metabolism or uptake, but the caveats described make it difficult to do so. Alternatively, we have to question what an elevation in blood pressure in response to these compounds means if 5-HT is a purported antihypertensive agent.

B. Clinical Situations

1. Circulatory Shock.

The involvement of 5-HT in shock, a condition of near cardiovascular collapse because of significant blood loss, is interesting and complicated. With the loss of blood, the body loses a significant portion of its 5-HT, because platelets are the primary source of circulating 5-HT. Shock is a 5-HT-reduced condition. Scrogin (2003) and Tiniakov et al. (2007) have done elegant work in demonstrating that activation of the 5-HT1A receptor centrally can reverse the profound reduction in total peripheral resistance in shock that is due to hemorrhage and can rescue blood pressure. The 5-HT1A receptor has markedly complex roles in blood pressure regulation, as outlined in Table 3. Depending on the site of injection and the dose, 8-OH-DPAT (which has significant affinity for 5-HT1 and 5-HT7 receptors) can either elevate or reduce blood pressure in the rat, the species in which the greatest amount of work has been performed. The 5-HT1A receptor clearly serves multiple purposes, but in shock, the 5-HT1A receptor supports blood pressure.

2. Orthostatic Hypotension.

Orthostatic hypotension (postural hypotension) is defined clinically as a ≥20 mm Hg decrease of systolic blood pressure or ≥10 mm Hg decrease of diastolic blood pressure upon standing (Medow et al., 2008). The dizziness experienced, associated with fainting, can limit the use of these medications, which occurs with a majority of the TCAs. The association of orthostatic hypotension with multiple types of serotonergic inhibitors raises the question of the involvement of 5-HT in this event. We cannot explain how SSRIs such as fluoxetine are used to treat orthostatic hypotension whereas TCAs cause hypotension, but it is logical to believe 5-HT is not at the root cause of either event.

3. Serotonin Syndrome.

More aptly named serotonin toxicity or serotonin toxidrome, this situation occurs with an overdose or interaction of drugs that promote an elevation of 5-HT (Skop et al., 1994; Mason et al., 2000; Boyer and Shannon, 2005). This is best recognized with ingestion of foods rich in tryptophan taken with MAO inhibitors, TCAs, SSRIs, SNRIs, or sometimes a combination of these drugs. This has also been observed with alternative, nonregulated therapies, such as St. John's wort and tryptophan. 5-HT toxicity presents with a multitude of symptoms that include central, gastrointestinal, muscular, autonomic, and cardiovascular symptoms. Of these, an elevation in temperature is notable. Hypertension is also associated with 5-HT toxicity. Obviously, 5-HT has a multitude of physiological effects that make its ability to reduce blood pressure, as we observed in infusions, obfuscated by other activated pathways.

4. Hypertension.

The role played by 5-HT in hypertension is still unclear. With the discovery of 5-HT in blood, it made sense that 5-HT would be regarded as a pathological factor in hypertension, promoting an increase in total peripheral resistance. Indeed, arteries isolated from both human and experimental models of hypertension show a classic hyper-reactivity to 5-HT that would support this activity. However, the action of 5-HT in vivo is far more complex. In the human, free circulating 5-HT is elevated, and the uptake of 5-HT appears to be impaired in the platelet of the hypertensive human such that the platelet appears “activated” (Kamal et al., 1984; Persson et al., 1988; Fetkovska et al., 1990a,b; Carrasco et al., 1998; Brenner et al., 2007).

By contrast, in pregnancy-induced hypertension, platelet content of 5-HT is increased and release of 5-HT is reduced (Gujrati et al., 1994) or shown to be unchanged (Jelen et al., 1979). Filshie et al. (1992) found that urinary 5-HIAA was not elevated in women with pregnancy-induced hypertension compared with women with normal pregnancy. Thus, there is no clear-cut association between platelet 5-HT content and blood pressure. The association of a reduced level of circulating 5-HT with elevated blood pressure in the human was suggested by Topsakal et al. (2009). In patients with hypertension who do not show the normal decrease in nocturnal blood pressure (“nondippers”), thrombocyte 5-HT levels were lower compared with “dippers” and control dubjects. We made an attempt to address this question of whether there is a correlative relationship between circulating 5-HT levels (free and platelet-bound) and blood pressure. For this purpose, we performed a regressive correlation of the free plasma and platelet-rich measures of 5-HT circulating in the rodent, along with telemetric measures of systolic blood pressure taken at the end of a 1-week infusion of 5-HT [Alzet miniosmotic pump (25 μg · kg−1 · min−1)]. These results were retrieved from our published articles and represent only male rats—both normotensive and hypertensive—that received 5-HT (Diaz et al., 2008; Davis et al., 2011; 25 μg · kg−1 · min−1 for 1 week). The 5-HT values in Fig. 8 represent the 5-HT concentration measured on the last day of infusion. Using these values, there is no linear relationship between either platelet-poor (free) or platelet-rich (plasma) and mean arterial blood pressure in the rat. As such, it is difficult to describe changes in circulating 5-HT as causal or a result of elevated blood pressure.

Fig. 8.

Fig. 8.

Correlation between platelet poor plasma (x-axis, left) and rich plasma (right) with mean arterial blood pressure (telemetric) taken from conscious rats at the end of a 1-week infusion of 5-HT.

The pathogenic effect of 5-HT was initially supported by studies reported decades ago that showed the ability of the 5-HT2A/2C receptor antagonist ketanserin to reduce blood pressure in the human and in experimental models of hypertension (Nelson et al., 1987; Liu et al., 1991; Azzadin et al., 1995; Krygicz et al., 1996; van Schie et al., 2002), but this effect was later attributed to α-adrenergic receptor blockade (Vanhoutte 1991; Wenting et al., 1982; Cohen et al., 1983a; Vanhoutte et al., 1988; van Zwieten et al., 1992). Other 5-HT2A receptor antagonists such as ritanserin and 6-methyl-1-(1-methylethyl)ergoline-8â-carboxylic acid 2-hydroxy-1-methylpropyl ester (LY53857) have not proven effective in reducing blood pressure (Cohen et al., 1983b; Gradin et al., 1985; Dalton et al., 1986; Nelson et al., 1987; Frishman et al., 1988; Stott et al., 1988; Docherty, 1989), making the effect of 5-HT on blood pressure equivocal. In the human, one report suggests that polymorphism of the 5-HT2A receptor is associated with elevation of blood pressure (Halder et al., 2007), but the mechanism of contribution of this polymorphism to blood pressure is not understood.

VII. Conclusions and Outstanding Questions

The role of 5-HT in modifying blood pressure is complex. We present both in vitro and in vivo evidence that argues 5-HT does have effects on blood pressure, many times opposite in nature in in vitro versus in vivo settings. These studies have raised a number of important questions that are the basis for future studies:

  • What is the mechanism of 5-HT-induced long-term depression of blood pressure? Time and again, it has been demonstrated that 5-HT causes a long-term decrease in blood pressure. This decrease seems to be completely dependent on 5-HT stimulation of NOS activity, because the NOS inhibitor l-NNA abolishes the effect of 5-HT. Where in the body does 5-HT stimulate NOS? The vascular endothelial cell and neuron are the best candidates.

  • Does 5-HT cross the blood-brain barrier? This question is important to answer given the long-held belief that 5-HT does not cross the blood-brain barrier, a belief held despite data, though decades old, that suggest the opposite. This knowledge would help us include or eliminate the central nervous system as a target for 5-HT in reducing blood pressure during prolonged exposure.

  • How are temperature and blood pressure linked by 5-HT? In several studies, 5-HT decreases blood pressure while causing heat loss. The ability of 5-HT to modify temperature itself, but also change temperature through cutaneous vasodilation, is interesting. Studies that investigate the primacy of temperature versus blood pressure homeostasis and whether 5-HT is a ‘connector’ to these two important physiological endpoints would be fascinating.

  • Does intracellular 5-HT modify cardiovascular function? The possibility that intracellular 5-HT can modify proteins has been raised. This is carried out by the enzyme transglutaminase to change the function of proteins such as the small G protein Rho. Paulmann et al. (2009) made the important observation that 5-HT regulates insulin secretion from pancreatic β cells by serotonylating (e.g., adding 5-HT to a protein covalently) GTPases. All the work cited above considers the biological activities of 5-HT as being exerted external to the cell through classic receptor stimulation. This particular article points to the idea that we may need to reconsider whether 5-HT also exerts biological effects internal to the cell.

Acknowledgments

This work was supported by the National Institutes of Health National Heart, Lung, and Blood Institute [Grant HL081115] (to S.W.W.). We thank Christopher McKee for the illustration in Fig. 2.

Authorship Contributions

Wrote or contributed to the writing of the manuscript: Watts, Morrison, Davis and Barman.

1
Abbreviations:
5-CT
5-carboxamidotryptamine
5-HIAA
5-hydroxyindole acetic acid
5-HT
5-hydroxytryptamine (serotonin)
5-HTP
5-hydroxytryptophan
8-OH-DPAT
8-hydroxy-2-(di-n-propylamino)tetralin
AVA
arteriovenous anastomoses
CNS
central nervous system
CV
cardiovascular
DOCA
deoxycorticosterone salt
GR127935
N-[4-methoxy-3-(4-methyl-1-piperazinyl)phenyl]-2′-methyl-4′-(5-methyl-1,2,4-oxadiazol-3-yl)-1–1′-biphenyl-4-carboxamide
HR
heart rate
IDO
indoleamine dioxygenase
IML
intermediolateral nucleus
KO
knockout
l-NNA
NG-nitro-L-arginine
MAO
monoamine oxidase
NE
norepinephrine
NOS
nitric-oxide synthase
NTS
nucleus of the tractus solitarius
PCPA
parachlorophenylalanine
PCPAME
parachlorophenylalanine methyl ester
RVLM
rostral ventrolateral medulla
SERT
serotonin transporter
SHR
spontaneously hypertensive rat
SNRI
serotonin/norepinephrine-reuptake inhibitor
SSRI
selective serotonin-reuptake inhibitor
TCA
tricyclic antidepressant
TPH
tryptophan hydroxylase
WT
wild type.

References

  1. Abassi ZA, Binah O, Youdim MB. (2004) Cardiovascular activity of rasagiline, a selective and potent inhibitor of mitochondrial monoamine oxidase B: comparison with selegiline. Br J Pharmacol 143:371–378 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Abumaria N, Ribic A, Anacker C, Fuchs E, Flügge G. (2008) Stress upregulates TPH1 but not TPH2 mRNA in the rat dorsal raphe nucleus: identification of two TPH2 mRNA splice variants. Cell Mol Neurobiol 28:331–342 [DOI] [PubMed] [Google Scholar]
  3. Afergan E, Epstein H, Dahan R, Koroukhov N, Rohekar K, Danenberg HD, Golomb G. (2008) Delivery of serotonin to the brain by monocytes following phagocytosis of liposomes. J Control Release 132:84–90 [DOI] [PubMed] [Google Scholar]
  4. Alenina N, Kikic D, Todiras M, Mosienko V, Qadri F, Plehm R, Boyé P, Vilianovitch L, Sohr R, Tenner K, et al. (2009) Growth retardation and altered autonomic control in mice lacking brain serotonin. Proc Natl Acad Sci USA 106:10332–10337 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Alexandrino-Silva C, Nadalini Mauá FH, de Andrade AG, de Toledo Ferraz Alves TC. (2008) Hypotension caused by therapeutic doses of venlafaxine: case report and proposed pathophysiological mechanisms. J Psychopharmacol 22:214–216 [DOI] [PubMed] [Google Scholar]
  6. Alkadhi KA, Alzoubi KH, Aleisa AM, Tanner FL, Nimer AS. (2005) Psychosocial stress-induced hypertension results from in vivo expression of long-term potentiation in rat sympathetic ganglia. Neurobiol Dis 20:849–857 [DOI] [PubMed] [Google Scholar]
  7. Alsip NL, Harris PD. (1992) Serotonin-induced dilation of small arterioles is not mediated via endothelium-derived relaxing factor in skeletal muscle. Eur J Pharmacol 229:117–124 [DOI] [PubMed] [Google Scholar]
  8. Alsip NL, Schuschke DA, Miller FN. (1996) Microvascular responses in the skeletal muscle of the diabetic rat. J Lab Clin Med 128:429–437 [DOI] [PubMed] [Google Scholar]
  9. Althaus JS, Beckman JJ, Miller ED., Jr (1985) Central serotonin depletion: effect on blood pressure during anesthesia. Anesth Analg 64:1163–1170 [PubMed] [Google Scholar]
  10. Amsterdam JD, Garcia-Espana F, Fawcett J, Quitkin FM, Reimherr FW, Rosenbaum JF, Beasley C. (1999) Blood pressure changes during short-term fluoxetine treatment. J Clin Psychopharmacol 19:9–14 [DOI] [PubMed] [Google Scholar]
  11. Anderson IK, Martin GR, Ramage AG. (1992) Central administration of 5-HT activates 5-HT1A receptors to cause sympathoexcitation and 5-HT2/5-HT1C receptors to release vasopressin in anaesthetized rats. Br J Pharmacol 107:1020–1028 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Anderson IK, Ramage AG, Gardiner SM. (1996) Cardiovascular effects of serotonin and DP-5-CT in conscious Long-Evans and Brattleboro rats. Am J Physiol 271:R455–R463 [DOI] [PubMed] [Google Scholar]
  13. Arima T, Ohshima Y, Mizuno T, Kitamura Y, Segawa T, Nomura Y. (1996) Cyclic GMP elevation by 5-hydroxytryptamine is due to nitric oxide derived from endogenous nitrosothiol in NG108–15 cells. Biochem Biophys Res Commun 227:473–478 [DOI] [PubMed] [Google Scholar]
  14. Asher EF, Alsip NL, Zhang PY, Harris PD. (1991) Halothane attenuates small arteriole vasodilation to serotonin (5-HT) in skeletal muscle. Anesth Analg 73:583–589 [DOI] [PubMed] [Google Scholar]
  15. Ashkenazi R, Finberg JP, Youdim MB. (1983) Effects of LM 5008, a selective inhibitor of 5-hydroxytryptamine uptake, on blood pressure and responses to sympathomimetic amines. Br J Pharmacol 79:915–922 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Azmitia EC. (2001) Modern views on an ancient chemical: serotonin effects on cell proliferation, maturation, and apoptosis. Brain Res Bull 56:413–424 [DOI] [PubMed] [Google Scholar]
  17. Azzadin A, Mysliwiec J, Wollny T, Mysliwiec M, Buczko W. (1995) Serotonin is involved in the pathogenesis of hypertension developing during erythropoietin treatment in uremic rats. Thromb Res 77:217–224 [DOI] [PubMed] [Google Scholar]
  18. Bagale SM, Brown AS, Gonzalez MM, Vitores A, Micotto TL, Kumar NS, Hentall ID, Wilson JN. (2011) Fluorescent reporters of monoamine transporter distribution and function. Bioorg Med Chem Lett 21:7387–7391 [DOI] [PubMed] [Google Scholar]
  19. Bagdy G, Calogero AE, Murphy DL, Szemeredi K. (1989) Serotonin agonists cause parallel activation of the sympathoadrenomedullary system and the hypothalamo-pituitary-adrenocortical axis in conscious rats. Endocrinology 125:2664–2669 [DOI] [PubMed] [Google Scholar]
  20. Bailey RK, Mallinckrodt CH, Wohlreich MM, Watkin JG, Plewes JM. (2006) Duloxetine in the treatment of major depressive disorder: comparisons of safety and efficacy. J Natl Med Assoc 98:437–447 [PMC free article] [PubMed] [Google Scholar]
  21. Banes AK, Watts SW. (2002) Upregulation of arterial serotonin 1B and 2B receptors in deoxycorticosterone acetate-salt hypertension. Hypertension 39:394–398 [DOI] [PubMed] [Google Scholar]
  22. Banes AK, Watts SW. (2003) Arterial expression of 5-HT2B and 5-HT1B receptors during development of DOCA-salt hypertension. BMC Pharmacology 3:12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Barman SM, Gebber GL. (2000) “Rapid” rhythmic discharges of sympathetic nerves: sources, mechanisms of generation, and physiological relevance. J Biol Rhythms 15:365–379 [DOI] [PubMed] [Google Scholar]
  24. Barnes NM, Sharp T. (1999) A review of central 5-HT receptors and their function. Neuropharmacology 38:1083–1152 [DOI] [PubMed] [Google Scholar]
  25. Barney CC, Threatte RM, Kikta DC, Fregly MJ. (1981) Effects of serotonin and L-5-hydroxytryptophan on plasma renin activity in rats. Pharmacol Biochem Behav 14:895–900 [DOI] [PubMed] [Google Scholar]
  26. Baron A, Riesselmann A, Fregly MJ. (1991) Reduction in the elevated blood pressure of Dahl salt-sensitive rats treated chronically with L-5-hydroxytryptophan. Pharmacology 42:15–22 [DOI] [PubMed] [Google Scholar]
  27. Bax WA, Renzenbrink GJ, Van Heuven-Nolsen D, Thijssen EJ, Bos E, Saxena PR. (1993) 5-HT receptors mediating contractions of the isolated human coronary artery. Eur J Pharmacol 239:203–210 [DOI] [PubMed] [Google Scholar]
  28. Bellou A, Lambert H, Gillois P, Montémont C, Gerard P, Vauthier E, Sainte-Laudy J, Longrois D, Guéant JL, Mallié JP. (2003) Constitutive nitric oxide synthase inhibition combined with histamine and serotonin receptor blockade improves the initial ovalbumin-induced arterial hypotension but decreases the survival time in brown norway rats anaphylactic shock. Shock 19:71–78 [DOI] [PubMed] [Google Scholar]
  29. Berger M, Gray JA, Roth BL. (2009) The expanded biology of serotonin. Ann Rev Med 60:355–366 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Bianchi P, Pimentel DR, Murphy MP, Colucci WS, Parini A. (2005) A new hypertrophic mechanism of serotonin in cardiac myocytes: receptor-independent ROS generation. FASEB J 19:641–643 [DOI] [PubMed] [Google Scholar]
  31. Biondi ML, Marasini B, Bianchi E, Agostoni A. (1988) Plasma free and intraplatelet serotonin in patients with Raynaud's phenomenon. Int J Cardiol 19:335–339 [DOI] [PubMed] [Google Scholar]
  32. Birdsall TC. (1998) 5-Hydroxytryptophan: a clinically-effective serotonin precursor. Altern Med Rev 3:271–280 [PubMed] [Google Scholar]
  33. Birkenfeld AL, Schroeder C, Pischon T, Tank J, Luft FC, Sharma AM, Jordan J. (2005) Paradoxical effect of sibutramine on autonomic cardiovascular regulation in obese hypertensive patients–sibutramine and blood pressure. Clin Auton Res 15:200–206 [DOI] [PubMed] [Google Scholar]
  34. Blackmore WP. (1958) Effect of serotonin on renal hemodynamics and sodium excretion in the dog. Am J Physiol 193:639–643 [DOI] [PubMed] [Google Scholar]
  35. Blankfield RP. (2006) The thermoregulatory-vascular remodeling hypothesis: an explanation for essential hypertension. Med Hypotheses 66:1174–1178 [DOI] [PubMed] [Google Scholar]
  36. Blauw G, Bom AH, van Brummelen P, Camps J, Arndt JW, Verdouw PD, Chang PC, van Zwieten PA, Saxena PR. (1991) Effects of 5-hydroxytryptamine on capillary and arteriovenous anastomotic blood flow in the human hand and forearm and in the pig hind leg. J Cardiovasc Pharmacol 17:316–324 [DOI] [PubMed] [Google Scholar]
  37. Blauw GJ, van Brummelen P, Chang PC, Vermeij P, van Zwieten PA. (1988) Regional vascular effects of serotonin and ketanserin in young, healthy subjects. Hypertension 11:256–263 [DOI] [PubMed] [Google Scholar]
  38. Blessing WW, Seaman B. (2003) 5-Hydroxytryptamine(2A) receptors regulate sympathetic nerves constricting the cutaneous vascular bed in rabbits and rats. Neuroscience 117:939–948 [DOI] [PubMed] [Google Scholar]
  39. Bodelsson M, Arneklo-Nobin B, Törnebrandt K. (1990) Effect of cooling on smooth muscle response to 5-hydroxytryptamine in human hand veins. Acta Physiol Scand 140:331–339 [DOI] [PubMed] [Google Scholar]
  40. Bom AH, Verdouw PD, Saxena PR. (1989) Carotid haemodynamics in pigs during infusions of 8-OH-DPAT: reduction in arteriovenous shunting is mediated by 5-HT1-like receptors. Br J Pharmacol 96:125–132 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Borgdorff P, Fekkes D, Tangelder GJ. (2002) Hypotension caused by extracorporeal circulation: serotonin from pump-activated platelets triggers nitric oxide release. Circulation 106:2588–2593 [DOI] [PubMed] [Google Scholar]
  42. Borgdorff P, Tangelder GJ. (2006) Pump-induced platelet aggregation with subsequent hypotension: its mechanism and prevention with clopidogrel. J Thorac Cardiovasc Surg 131:813–821 [DOI] [PubMed] [Google Scholar]
  43. Borton M, Neligan M, Wood F, Dervan P, Goggins K, Docherty JR. (1990) Contractions to 5-hydroxytryptamine in human coronary artery and human saphenous vein. Br J Clin Pharmacol 30 (Suppl 1):107S–108S [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Boston PC, Hodgson WC. (1997) Changes in reactivity towards 5-hydroxytryptamine in the renal vasculature of the diabetic spontaneously hypertensive rat. J Hypertens 15:769–774 [DOI] [PubMed] [Google Scholar]
  45. Bouchelet I, Case B, Olivier A, Hamel E. (2000) No contractile effect for 5-HT1D and 5-HT1F receptor agonists in human and bovine cerebral arteries: similarity with human coronary artery. Br J Pharmacol 129:501–508 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Boyer EW, Shannon M. (2005) The serotonin syndrome. N Engl J Med 352:1112–1120 [DOI] [PubMed] [Google Scholar]
  47. Brand T, Anderson GM. (2011) The measurement of platelet-poor plasma serotonin: a systematic review of prior reports and recommendations for improved analysis. Clin Chem 57:1376–1386 [DOI] [PubMed] [Google Scholar]
  48. Brenner B, Harney JT, Ahmed BA, Jeffus BC, Unal R, Mehta JL, Kilic F. (2007) Plasma serotonin levels and the platelet serotonin transporter. J Neurochem 102:206–215 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Breuer J, Georgaraki A, Sieverding L, Baden W, Apitz J. (1996) Increased turnover of serotonin in children with pulmonary hypertension secondary to congenital heart disease. Pediatr Cardiol 17:214–219 [DOI] [PubMed] [Google Scholar]
  50. Brownfield MS, Poff BC, Holzwarth MA. (1985) Ultrastructural immunocytochemical co-localization of serotonin and PNMT in adrenal medullary vesicles. Histochemistry 83:41–46 [DOI] [PubMed] [Google Scholar]
  51. Brudeli B, Moltzau LR, Andressen KW, Krobert KA, Klaveness J, Levy FO. (2010) Synthesis and pharmacological properties of novel hydrophilic 5-HT4 receptor antagonists. Bioorg Med Chem 18:8600–8613 [DOI] [PubMed] [Google Scholar]
  52. Bruning TA, Chang PC, Blauw GJ, Vermeij P, van Zwieten PA. (1993) Serotonin-induced vasodilatation in the human forearm is mediated by the “nitric oxide-pathway”: no evidence for involvement of the 5-HT3-receptor. J Cardiovasc Pharmacol 22:44–51 [DOI] [PubMed] [Google Scholar]
  53. Bruning TA, van Zwieten PA, Blauw GJ, Chang PC. (1994) No functional involvement of 5-hydroxytryptamine1A receptors in nitric oxide-dependent dilatation caused by serotonin in the human forearm vascular bed. J Cardiovasc Pharmacol 24:454–461 [DOI] [PubMed] [Google Scholar]
  54. Brust P, Friedrich A, Krizbai IA, Bergmann R, Roux F, Ganapathy V, Johannsen B. (2000) Functional expression of the serotonin transporter in immortalized rat brain microvessel endothelial cells. J Neurochem 74:1241–1248 [DOI] [PubMed] [Google Scholar]
  55. Buckingham RE, Hamilton TC, Moore RA. (1976) Prolonged effects of p-chlorophenylalanine on the blood pressure of conscious normotensive and DOCA/saline hypertensive rats. Br J Pharmacol 56:69–75 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Bulat M, Supek Z. (1968) Passage of 5-hydroxytryptamine through the blood-brain barrier, its metabolism in the brain and elimination of 5-hydroxyindoleacetic acid from the brain tissue. J Neurochem 15:383–389 [DOI] [PubMed] [Google Scholar]
  57. Cai SR, Wang LC, Kong DH, Huang ZX, Ma RC. (1999) Substance P- and 5-hydroxytryptamine-mediated depolarization in sympathetic ganglion neurons. Acta Physiol Sinica 51:585–587 [PubMed] [Google Scholar]
  58. Calama E, García M, Jarque MJ, Morán A, Martín ML, San Román L. (2003) 5-Hydroxytryptamine-induced vasodilator responses in the hindquarters of the anaesthetized rat, involve beta2 adrenoceptors. J Pharm Pharmacol 55:1371–1378 [DOI] [PubMed] [Google Scholar]
  59. Calama E, Ortíz de Urbina AV, Morán A, Martín ML, San Román L. (2005) Effect of 5-hydroxytryptamine on neurogenic vasoconstriction in the isolated, autoperfused hindquarters of the rat. Clin Exp Pharmacol Physiol 32:894–900 [DOI] [PubMed] [Google Scholar]
  60. Callera JC, Bonagamba LG, Sévoz C, Laguzzi R, Machado BH. (1997) Cardiovascular effects of microinjection of low doses of serotonin into the NTS of unanesthetized rats. Am J Physiol 272:R1135–R1142 [DOI] [PubMed] [Google Scholar]
  61. Callera JC, Colombari E, De Luca LA, Jr, Menani JV. (2005) The bradycardic and hypotensive responses to serotonin are reduce by activation of GABAA receptors in the nucleus tractus solitarius of awake rats. Braz J Med Biol Res 38:1123–1131 [DOI] [PubMed] [Google Scholar]
  62. Carlson LA, Ekelund LG, Orö L. (1967) Metabolic and cardio-vascular effects of serotonin. Life Sci 6:261–271 [DOI] [PubMed] [Google Scholar]
  63. Carlsson A, Wong DT. (1997) Correction: a note on the discovery of selective serotonin reuptake inhibitors. Life Sci 61:1203. [DOI] [PubMed] [Google Scholar]
  64. Caron J, Libersa C, Hazard JR, Lacroix D, Facq E, Guedon-Moreau L, Fautrez V, Solles A, Puozzo C, Kacet S. (1993) Acute electrophysiological effects of intravenous milnacipran, a new antidepressant agent. Eur Neuropsychopharmacol 3:493–500 [DOI] [PubMed] [Google Scholar]
  65. Carrasco G, Cruz MA, Gallardo V, Miguel P, Lagos M, González C. (1998) Plasma and platelet concentration and platelet uptake of serotonin in normal and pre-eclamptic pregnancies. Life Sci 62:1323–1332 [DOI] [PubMed] [Google Scholar]
  66. Cartier D, Jégou S, Parmentier F, Lihrmann I, Louiset E, Kuhn JM, Bastard C, Plouin PF, Godin M, Vaudry H, et al. (2005) Expression profile of serotonin4 (5-HT4) receptors in adrenocortical aldosterone-producing adenomas. Eur J Endocrinol 153:939–947 [DOI] [PubMed] [Google Scholar]
  67. Castejon AM, Paez X, Hernandez L, Cubeddu LX. (1999) Use of intravenous microdialysis to monitor changes in serotonin release and metabolism induced by cisplatin in cancer patients: comparative effects of granisetron and ondansetron. J Pharmacol Exp Ther 291:960–966 [PubMed] [Google Scholar]
  68. Cavero I, Lefèvre-Borg F, Roach AG. (1981) Effects of mianserin, desipramine and maprotiline on blood pressure responses evoked by acetylcholine, histamine and 5-hydroxytryptamine in rats. Br J Pharmacol 74:143–148 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Centurión D, Glusa E, Sánchez-López A, Valdivia LF, Saxena PR, Villalón CM. (2004) 5-HT7, but not 5-HT2B, receptors mediate hypotension in vagosympathectomized rats. Eur J Pharmacol 502:239–242 [DOI] [PubMed] [Google Scholar]
  70. Centurión D, Sánchez-López A, Ortiz MI, De Vries P, Saxena PR, Villalón CM. (2000) Mediation of 5-HT-induced internal carotid vasodilatation in GR127935- and ritanserin-pretreated dogs by 5-HT7 receptors. Naunyn Schmiedebergs Arch Pharmacol 362:169–176 [DOI] [PubMed] [Google Scholar]
  71. Chalon SA, Granier LA, Vandenhende FR, Bieck PR, Bymaster FP, Joliat MJ, Hirth C, Potter WZ. (2003) Duloxetine increases serotonin and norepinephrine availability in healthy subjects: a double-blind, controlled study. Neuropsychopharmacology 28:1685–1693 [DOI] [PubMed] [Google Scholar]
  72. Chanrion B, Mannoury la Cour C, Bertaso F, Lerner-Natoli M, Freissmuth M, Millan MJ, Bockaert J, Marin P. (2007) Physical interaction between the serotonin transporter and neuronal nitric oxide synthase underlies reciprocal modulation of their activity. Proc Natl Acad Sci USA 104:8119–8124 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Chen JJ, Li Z, Pan H, Murphy DL, Tamir H, Koepsell H, Gershon MD. (2001) Maintenance of serotonin in the intestinal mucosa and ganglia of mice that lack the high-affinity serotonin transporter: abnormal intestinal motility and the expression of cation transporters. J Neurosci 21:6348–6361 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Chester AH, Martin GR, Bodelsson M, Arneklo-Nobin B, Tadjkarimi S, Tornebrandt K, Yacoub MH. (1990) 5-Hydroxytryptamine receptor profile in healthy and diseased human epicardial coronary arteries. Cardiovasc Res 24:932–937 [DOI] [PubMed] [Google Scholar]
  75. Christensen P, Thomsen HY, Pedersen OL, Gram LF, Kragh-Sørensen P. (1985) Orthostatic side effects of clomipramine and citalopram during treatment for depression. Psychopharmacology 86:383–385 [DOI] [PubMed] [Google Scholar]
  76. Chudasama HP, Bhatt PA. (2009) Evaluation of anti-obesity activity of duloxetine in comparison with sibutramine along with its anti-depressant activity: an experimental study in obese rats. Can J Physiol Pharmacol 87:900–907 [DOI] [PubMed] [Google Scholar]
  77. Cohen ML, Fuller RW, Kurz KD. (1983a) Evidence that blood pressure reduction by serotonin antagonists is related to alpha receptor blockade in spontaneously hypertensive rats. Hypertension 5:676–681 [DOI] [PubMed] [Google Scholar]
  78. Cohen ML, Fuller RW, Kurz KD. (1983b) LY53857, a selective and potent serotonergic (5-HT2) receptor antagonist, does not lower blood pressure in the spontaneously hypertensive rat. J Pharmacol Exp Ther 227:327–332 [PubMed] [Google Scholar]
  79. Cohen RA. (1986) Contractions of isolated canine coronary arteries resistant to S2-serotonergic blockade. J Pharmacol Exp Ther 237:548–552 [PubMed] [Google Scholar]
  80. Comet MA, Bernard JF, Hamon M, Laguzzi R, Sévoz-Couche C. (2007) Activation of nucleus tractus solitarius 5-HT2A but not other 5-HT2 receptor subtypes inhibits the sympathetic activity in rats. Eur J Neurosci 26:345–354 [DOI] [PubMed] [Google Scholar]
  81. Contesse V, Lenglet S, Grumolato L, Anouar Y, Lihrmann I, Lefebvre H, Delarue C, Vaudry H. (1999) Pharmacological and molecular characterization of 5-hydroxytryptamine(7) receptors in the rat adrenal gland. Mol Pharmacol 56:552–561 [DOI] [PubMed] [Google Scholar]
  82. Côté F, Fligny C, Fromes Y, Mallet J, Vodjdani G. (2004) Recent advances in understanding serotonin regulation of cardiovascular function. Trends Mol Med 10:232–238 [DOI] [PubMed] [Google Scholar]
  83. Côté F, Thévenot E, Fligny C, Fromes Y, Darmon M, Ripoche MA, Bayard E, Hanoun N, Saurini F, Lechat P, et al. (2003) Disruption of the nonneuronal tph1 gene demonstrates the importance of peripheral serotonin in cardiac function. Proc Natl Acad Sci USA 100:13525–13530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Csaba G, Sudár F. (1978) Localization of radioactively labelled serotonin in the nucleus of adrenal medulla cells. Acta Anat 100:237–240 [DOI] [PubMed] [Google Scholar]
  85. Cummings SA, Groszmann RJ, Kaumann AJ. (1986) Hypersensitivity of mesenteric veins to 5-hydroxytryptamine- and ketanserin-induced reduction of portal pressure in portal hypertensive rats. Br J Pharmacol 89:501–513 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Cushing DJ, Baez M, Kursar JD, Schenck K, Cohen ML. (1994) Serotonin-induced contraction in canine coronary artery and saphenous vein: role of a 5-HT1D-like receptor. Life Sci 54:1671–1680 [DOI] [PubMed] [Google Scholar]
  87. Cushing DJ, Cohen ML. (1992a) Serotonin-induced relaxation in canine coronary artery smooth muscle. J Pharmacol Exp Ther 263:123–129 [PubMed] [Google Scholar]
  88. Cushing DJ, Cohen ML. (1992b) Comparison of the serotonin receptors that mediate smooth muscle contraction in canine and porcine coronary artery. J Pharmacol Exp Ther 261:856–862 [PubMed] [Google Scholar]
  89. Cushing DJ, Cohen ML. (1993) Serotonin-induced contraction in porcine coronary artery: use of ergolines to support vascular 5-hydroxytryptamine2-receptor heterogeneity. J Pharmacol Exp Ther 264:193–200 [PubMed] [Google Scholar]
  90. Cushing DJ, Zgombick JM, Nelson DL, Cohen ML. (1996) LY215840, a high-affinity 5-HT7 receptor ligand, blocks serotonin-induced relaxation in canine coronary artery. J Pharmacol Exp Ther 277:1560–1566 [PubMed] [Google Scholar]
  91. Dabiré H, Laubie M, Schmitt H. (1990) Hypotensive effects of 5-HT1A receptor agonists on the ventrolateral medullary pressor area in dogs. J Cardiovasc Pharmacol 15 (Suppl 7):S61–S67 [PubMed] [Google Scholar]
  92. Dalton DW, Feniuk W, Humphrey PP. (1986) An investigation into the mechanisms of the cardiovascular effects of 5-hydroxytryptamine in conscious normotensive and DOCA-salt hypertensive rats. J Auton Pharmacol 6:219–228 [DOI] [PubMed] [Google Scholar]
  93. Das YT, Bagchi M, Bagchi D, Preuss HG. (2004) Safety of 5-hydroxy-l-tryptophan. Toxicol Lett 150:111–122 [DOI] [PubMed] [Google Scholar]
  94. Davis RP, Linder AE, Watts SW. (2011) Lack of the serotonin transporter (SERT) reduces the ability of 5-hydroxytryptamine to lower blood pressure. Naunyn Schmiedebergs Arch Pharmacol 383:543–546 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Daws LC. (2009) Unfaithful neurotransmitter transporters: focus on serotonin uptake and implications for antidepressant efficacy. Pharmacol Ther 121:89–99 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. de Lemos HP, Jr, Atallah AN, de Lemos AL. (2008) Can sibutramine alter systemic blood pressure in obese patients? Systematic review and meta-analysis. Sao Paulo Med J 126:342–346 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. De Vries P, De Visser PA, Heiligers JP, Villalón CM, Saxena PR. (1999) Changes in systemic and regional haemodynamics during 5-HT7 receptor-mediated depressor responses in rats. Naunyn-Schmiedebergs Arch Pharmacol 359:331–338 [DOI] [PubMed] [Google Scholar]
  98. Dedeoğlu A, Fisher LA. (1991) Central nervous actions of serotonin and a serotonin1A receptor agonist: cardiovascular excitation at low doses. J Pharmacol Exp Ther 257:425–432 [PubMed] [Google Scholar]
  99. Delarue C, Becquet D, Idres S, Hery F, Vaudry H. (1992) Serotonin synthesis in adrenochromaffin cells. Neuroscience 46:495–500 [DOI] [PubMed] [Google Scholar]
  100. Derby MA, Zhang L, Chappell JC, Gonzales CR, Callaghan JT, Leibowitz M, Ereshefsky L, Hoelscher D, Leese PT, Mitchell MI. (2007) The effects of a supratherapeutic doses of duloxetine on blood pressure and pulse rate. J Cardiovasc Pharmacol 49:384–393 [DOI] [PubMed] [Google Scholar]
  101. Diaz J, Ni W, Thompson J, King A, Fink GD, Watts SW. (2008) 5-Hydroxytryptamine lowers blood pressure in normotensive and hypertensive rats. J Pharmacol Exp Ther 325:1031–1038 [DOI] [PubMed] [Google Scholar]
  102. Ding XR, Stier CT, Jr, Itskovitz HD. (1989) Serotonin and 5-hydroxytryptophan on blood pressure and renal blood flow in anesthetized rats. Am J Med Sci 297:290–293 [DOI] [PubMed] [Google Scholar]
  103. Ding YA, Chou TC, Huan R. (1994) Are platelet cytosolic free calcium, serotonin concentration and blood viscosity different between hypertensive and normotensive subjects? Cardiology 85:76–81 [DOI] [PubMed] [Google Scholar]
  104. Dobbins DE, Adamski SW, Lokhandwala MF, Grega GJ. (1983) Evidence that serotonin receptors mediate the cutaneous vasoconstriction produced by 5-hydroxytryptamine in canine forelimbs. Circ Res 53:473–481 [DOI] [PubMed] [Google Scholar]
  105. Docherty JR. (1988) Investigation of cardiovascular 5-hydroxytryptamine receptor subtypes in the rat. Naunyn Schmiedebergs Arch Pharmacol 337:1–8 [DOI] [PubMed] [Google Scholar]
  106. Docherty JR. (1989) An investigation of the effects of intravenous injection of the 5-hydroxytryptamine 2 receptor antagonists ketanserin and LY 53857 on blood pressure in anesthetized spontaneously hypertensive rats. J Pharmacol Exp Ther 248:736–740 [PubMed] [Google Scholar]
  107. Doggrell SA. (2003) The role of 5-HT on the cardiovascular and renal systems and the clinical potential of 5-HT modulation. Expert Opin Investig Drugs 12:805–823 [DOI] [PubMed] [Google Scholar]
  108. Dohi Y, Lüscher TF. (1991) Altered intra- and extraluminal effects of 5-hydroxytryptamine in hypertensive mesenteric resistance arteries: contribution of the endothelium and smooth muscle. J Cardiovasc Pharmacol 18:278–284 [DOI] [PubMed] [Google Scholar]
  109. Dorigo P, Belluco P, Corradi L, Gaion RM, Maragno I. (1992) Vasodilatory activity of etozoline in rat and guinea-pig isolated aorta: study of the mechanism of action. Arch Int Pharmacodyn Ther 315:63–78 [PubMed] [Google Scholar]
  110. Dun NJ, Karczmar AG. (1981) Evidence for a presynaptic inhibitory action of 5-hydroxytryptamine in a mammalian sympathetic ganglion. J Pharmacol Exp Ther 217:714–718 [PubMed] [Google Scholar]
  111. Dun NJ, Kiraly M, Ma RC. (1984) Evidence for a serotonin-mediated slow excitatory potential in the guinea-pig coeliac ganglia. J Physiol 351:61–76 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Echizen H, Freed CR. (1982) Long-term infusion of L-5-hydroxytryptophan increases brain serotonin turnover and decreases blood pressure in normotensive rats. J Pharmacol Exp Ther 220:579–584 [PubMed] [Google Scholar]
  113. Edvinsson L, Uddman E, Wackenfors A, Davenport A, Longmore J, Malmsjö M. (2005) Triptan-induced contractile (5-HT1B receptor) responses in human cerebral and coronary arteries: relationship to clinical effect. Clin Sci 109:335–342 [DOI] [PubMed] [Google Scholar]
  114. Eglen RM, Perkins LA, Walsh LK, Whiting RL. (1992) Agonist action of indole derivatives at 5-HT1-like, 5-HT3, and 5HT4 receptors in vitro. J Auton Pharmacol 12:321–333 [DOI] [PubMed] [Google Scholar]
  115. Ellis ES, Byrne C, Murphy OE, Tilford NS, Baxter GS. (1995) Mediation by 5-hydroxytryptamine2B receptors of endothelium-dependent relaxation in rat jugular vein. Br J Pharmacol 114:400–404 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Ellwood AJ, Curtis MJ. (1997) Involvement of 5-HT(1B/1D) and 5-HT2A receptors in 5-HT-induced contraction of endothelium-denuded rabbit epicardial coronary arteries. Br J Pharmacol 122:875–884 [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Erspamer V, Asero B. (1952) Identification of enteramine, the specific hormone of the enterochromaffin cell system, as 5-hydroxytryptamine. Nature 169:800–801 [DOI] [PubMed] [Google Scholar]
  118. Evans RG, Haynes JM, Ludbrook J. (1993) Effects of 5-HT-receptor and alpha 2-adrenoceptor ligands on the haemodynamic response to acute central hypovolaemia in conscious rabbits. Br J Pharmacol 109:37–47 [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Fastier FN, Waal H. (1957) The antidiuretic action of 5-hydroxytryptamine in cats in relation to the production of certain chemoreflexes. Br J Pharmacol Chemother 12:484–488 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Feldman PD, Galiano FJ. (1995) Cardiovascular effects of serotonin in the nucleus of the solitary tract. Am J Physiol 269:R48–R56 [DOI] [PubMed] [Google Scholar]
  121. Feniuk W, Humphrey PP, Watts AD. (1979) Presynaptic inhibitory action of 5-hydroxytryptamine in dog isolated saphenous vein. Br J Pharmacol 67:247–254 [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Feniuk W, Humphrey PP, Perren MJ, Watts AD. (1985) A comparison of 5-hydroxytryptamine receptors mediating contraction in rabbit aorta and dog saphenous vein: evidence for different receptor types obtained by use of selective agonists and antagonists. Br J Pharmacol 86:697–704 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Ferreira HS, de Castro e Silva E, Cointeiro C, Oliveira E, Faustino TN, Fregoneze JB. (2004) Role of central 5-HT3 receptors in the control of blood pressure in stressed and non-stressed rats. Brain Res 1028:48–58 [DOI] [PubMed] [Google Scholar]
  124. Ferreira HS, Oliveira E, Faustino TN, Silva Ede C, Fregoneze JB. (2005) Effect of the activation of central 5-HT2C receptors by the 5-HT2C agonist mCPP on blood pressure and heart rate in rats. Brain Res 1040:64–72 [DOI] [PubMed] [Google Scholar]
  125. Fetkovska N, Amstein R, Ferracin F, Regenass M, Bühler FR, Pletscher A. (1990a) 5-Hydroxytryptamine kinetics and activation of blood platelets in patients with essential hypertension. Hypertension 15:267–273 [DOI] [PubMed] [Google Scholar]
  126. Fetkovska N, Pletscher A, Ferracin F, Amstein R, Buhler FR. (1990b) Impaired uptake of 5 hydroxytryptamine platelet in essential hypertension: clinical relevance. Cardiovasc Drugs Ther 4 (Suppl 1):105–109 [DOI] [PubMed] [Google Scholar]
  127. Feuerstein TJ. (2008) Presynaptic receptors for dopamine, histamine, and serotonin. Handb Exp Pharmacol 184:289–338 [DOI] [PubMed] [Google Scholar]
  128. Filshie GM, Maynard P, Hutter C, Cooper JC, Robinson G, Rubin P. (1992) Urinary 5-hydroxyindole acetate concentration in pregnancy induced hypertension. BMJ 304:1223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Finberg JP, Gross A, Bar-Am O, Friedman R, Loboda Y, Youdim MB. (2006) Cardiovascular responses to combined treatment with selective monoamine oxidase type B inhibitors and L-DOPA in the rat. Br J Pharmacol 149:647–656 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Flórez J, Armijo JA. (1974) Effect of central inhibition of the 1-amino acid decarboxylase on the hypotensive action of 5-HT precursors in cats. Eur J Pharmacol 26:108–110 [DOI] [PubMed] [Google Scholar]
  131. Fozard JR, Mir AK, Middlemiss DN. (1987) Cardiovascular response to 8-hydroxy-2-(di-n-propylamino) tetralin (8-OH-DPAT) in the rat: site of action and pharmacological analysis. J Cardiovasc Pharmacol 9:328–347 [DOI] [PubMed] [Google Scholar]
  132. Frandsen P, Nielsen AT. (1966) The antidiuretic effect of 5-hydroxytryptamine in the rat after subcutaneous injection of ‘physiological’ doses. Acta Pharmacol Toxicol (Copenh) 24:50–54 [DOI] [PubMed] [Google Scholar]
  133. Fregly MJ, Lockley OE, Sumners C. (1987a) Chronic treatment with l-5-hydroxytryptophan prevents the development of DOCA-salt-induced hypertension in rats. J Hypertens 5:621–628 [DOI] [PubMed] [Google Scholar]
  134. Fregly MJ, Lockley OE, van der Voort J, Sumners C, Henley WN. (1987b) Chronic dietary administration of tryptophan prevents the development of deoxycorticosterone acetate salt induced hypertension in rats. Can J Physiol Pharmacol 65:753–764 [DOI] [PubMed] [Google Scholar]
  135. Frishman WH, Okin S, Huberfeld S. (1988) Serotonin antagonism in the treatment of systemic hypertension: the role of ketanserin. Med Clin North Am 72:501–522 [DOI] [PubMed] [Google Scholar]
  136. Froldi G, Montopoli M, Zanetti M, Dorigo P, Caparrotta L. (2008) 5-HT1B receptor subtype and aging in rat resistance vessels. Pharmacology 81:70–78 [DOI] [PubMed] [Google Scholar]
  137. Froldi G, Nicoletti P, Caparrotta L, Ragazzi E. (2003) 5-HT receptors mediating contraction in the rat-tail artery. Adv Exp Med Biol 527:665–670 [DOI] [PubMed] [Google Scholar]
  138. Fuentes JA, Ordaz A, Neff NH. (1979) Central mediation of the antihypertensive effect of pargyline in spontaneously hypertensive rats. Eur J Pharmacol 57:21–27 [DOI] [PubMed] [Google Scholar]
  139. Fujiwara T, Chiba S. (1995) Augmented responses to 5-HT2-receptor-mediated vasoconstrictions in atherosclerotic rabbit common carotid arteries. J Cardiovasc Pharmacol 26:503–510 [DOI] [PubMed] [Google Scholar]
  140. Gaciong Z, Placha G. (2005) Efficacy and safety of sibutramine in 2225 subjects with cardiovascular risk factors: short-term, open-label, observational study. J Hum Hypertens 19:737–743 [DOI] [PubMed] [Google Scholar]
  141. Gaddum JH, Picarelli ZP. (1957) Two kinds of tryptamine receptor. Br J Pharmacol 12:323–328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Gale JD, Cowen T. (1988) The origin and distribution of 5-hydroxytryptamine-like immunoreactive nerve fibres to major mesenteric blood vessels of the rat. Neuroscience 24:1051–1059 [DOI] [PubMed] [Google Scholar]
  143. Galzin AM, Delahaye M, Hoornaert C, McCort G, O'Connor SE. (2000) Effects of SL 65.0472, a novel 5-HT receptor antagonist, on 5-HT receptor mediated vascular contraction. Eur J Pharmacol 404:361–368 [DOI] [PubMed] [Google Scholar]
  144. García M, Morán A, Luisa Martín M, Barthelmebs M, San Román L. (2006) The nitric oxide synthesis/pathway mediates the inhibitory serotoninergic responses of the pressor effect elicited by sympathetic stimulation in diabetic pithed rats. Eur J Pharmacol 537:126–134 [DOI] [PubMed] [Google Scholar]
  145. Gaw AJ, Wadsworth RM, Humphrey PP. (1990) Pharmacological characterisation of postjunctional 5-HT receptors in cerebral arteries from the sheep. Eur J Pharmacol 179:35–44 [DOI] [PubMed] [Google Scholar]
  146. Geerts IS, De Meyer GR, Bult H. (2000) Collar-induced elevation of mRNA and functional activity of 5-HT(1B) receptor in the rabbit carotid artery. Br J Pharmacol 131:1723–1731 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Gerges NZ, Aleisa AM, Alhaider AA, Alkadhi KA. (2002) Reduction of elevated arterial blood pressure in obese Zucker rats by inhibition of ganglionic long-term potentiation. Neuropharmacology 43:1070–1076 [DOI] [PubMed] [Google Scholar]
  148. Giardina EG, Johnson LL, Vita J, Bigger JT, Jr, Brem RF. (1985) Effect of imipramine and nortriptyline on left ventricular function and blood pressure in patients treated for arrhythmias. Am Heart J 109:992–998 [DOI] [PubMed] [Google Scholar]
  149. Gilbert MJ, Newberry NR. (1987) A 5-HT1-like receptor mediates a sympathetic ganglionic hyperpolarization. Eur J Pharmacol 144:385–388 [DOI] [PubMed] [Google Scholar]
  150. Gilmore B, Michael M. (2011) Treatment of acute migraine headache. Am Fam Physician 83:271–280 [PubMed] [Google Scholar]
  151. Glassman AH, Bigger JT, Jr, Giardina EV, Kantor SJ, Perel JM, Davies M. (1979) Clinical characteristics of imipramine-induced orthostatic hypotension. Lancet 1:468–472 [DOI] [PubMed] [Google Scholar]
  152. Glusa E, Müller-Schweinitzer E. (1993) Heterogeneity of 5-HT receptor subtypes in isolated human femoral and saphenous veins. Naunyn Schmiedebergs Arch Pharmacol 347:133–136 [DOI] [PubMed] [Google Scholar]
  153. Glusa E, Roos A. (1996) Endothelial 5-HT receptors mediate relaxation of porcine pulmonary arteries in response to ergotamine and dihydroergotamine. Br J Pharmacol 119:330–334 [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Göthert M, Molderings GJ, Fink K, Schlicker E. (1991) Heterogeneity of presynaptic serotonin receptors on sympathetic neurones in blood vessels. Blood Vessels 28:11–18 [DOI] [PubMed] [Google Scholar]
  155. Göthert M, Schlicker E, Kollecker P. (1986) Receptor-mediated effects of serotonin and 5-methoxytryptamine on noradrenaline release in the rat vena cava and in the heart of the pithed rat. Naunyn Schmiedebergs Arch Pharmacol 332:124–130 [DOI] [PubMed] [Google Scholar]
  156. Gradin K, Pettersson A, Hedner T, Persson B. (1985) Chronic 5-HT2 receptor blockade with ritanserin does not reduce blood pressure in the spontaneously hypertensive rat. J Neural Transm 64:145–149 [DOI] [PubMed] [Google Scholar]
  157. Grandaw PP, Purdy RE. (1996) Serotonin-induced vasoconstriction in rabbit femoral artery: mediation by both 5-HT2 serotonergic and alpha 1-adrenoceptors. J Cardiovasc Pharmacol 27:854–860 [DOI] [PubMed] [Google Scholar]
  158. Green AR. (2006) Neuropharmacology of 5-hydroxytryptamine. Br J Pharmacol 147:S145–S152 [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Grubb BP, Samoil D, Kosinski D, Wolfe D, Lorton M, Madu E. (1994) Fluoxetine hydrochloride for the treatment of severe refractory orthostatic hypotension. Am J Med 97:366–368 [DOI] [PubMed] [Google Scholar]
  160. Guelfi JD, Dreyfus JF, Pichot P. (1983) A double-blind controlled clinical trial comparing fluvoxamine with imipramine. Br J Clin Pharmacol 15:411S–417S [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Gujrati VR, Goyal A, Gaur SP, Singh N, Shanker K, Chandravati (1994) Relevance of platelet serotonergic mechanisms in pregnancy induced hypertension. Life Sci 55:327–335 [DOI] [PubMed] [Google Scholar]
  162. Gul H, Yildiz O, Simsek A, Balkan M, Ersoz N, Cetiner S, Isimer A, Sen D. (2003) Pharmacologic characterization of contractile serotonergic receptors in human isolated mesenteric artery. J Cardiovasc Pharmacol 41:307–315 [DOI] [PubMed] [Google Scholar]
  163. Guyenet PG. (2006) The sympathetic control of blood pressure. Nat Rev Neurosci 7:335–346 [DOI] [PubMed] [Google Scholar]
  164. Hafdi Z, Couette S, Comoy E, Prie D, Amiel C, Friedlander G. (1996) Locally formed 5-hydroxytryptamine stimulates phosphate transport in cultured opossum kidney cells and in rat kidney. Biochem J 320:615–621 [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Hajjo R, Grulke CM, Golbraikh A, Setola V, Huang XP, Roth BL, Tropsha A. (2010) Development, validation, and use of quantitative structure-activity relationship models of 5-hydroxytryptamine (2B) receptor ligands to identify novel receptor binders and putative valvulopathic compounds among common drugs. J Med Chem 53:7573–7586 [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Halder I, Muldoon MF, Ferrell RE, Manuck SB. (2007) Serotonin receptor 2A (HTR2A) gene polymorphisms are associated with blood pressure, central adiposity, and the metabolic syndrome. Metab Syndr Relat Disord 5:323–330 [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Hamel E, Grégoire L, Lau B. (1993) 5-HT1 receptors mediating contraction in bovine cerebral arteries: a model for human cerebrovascular ‘5-HT1D beta’ receptors. Eur J Pharmacol 242:75–82 [DOI] [PubMed] [Google Scholar]
  168. Hamel E, Robert JP, Young AR, MacKenzie ET. (1989) Pharmacological properties of the receptor(s) involved in the 5-hydroxytryptamine-induced contraction of the feline middle cerebral artery. J Pharmacol Exp Ther 249:879–889 [PubMed] [Google Scholar]
  169. Hansen MB, Arif F, Gregersen H, Bruusgaard H, Wallin L. (2008) Effect of serotonin on small intestinal contractility in healthy volunteers. Physiol Res 57:63–71 [DOI] [PubMed] [Google Scholar]
  170. Häppölä O. (1988) 5-Hydroxytryptamine-immunoreactive neurons and nerve fibers in the superior cervical ganglion of the rat. Neuroscience 27:301–307 [DOI] [PubMed] [Google Scholar]
  171. Hardcastle J, Hardcastle PT. (1999) Effects of 5-hydroxytryptamine on the transintestinal electrical activity and cardiovascular function of fawn-hooded rats in-vivo. J Pharm Pharmacol 51:1415–1421 [DOI] [PubMed] [Google Scholar]
  172. Hardebo JE, Owman C. (1980) Barrier mechanisms for neurotransmitter monoamines and their precursors at the blood-brain interface. Ann Neurol 8:1–31 [DOI] [PubMed] [Google Scholar]
  173. Haugen G, Mellembakken J, Stray-Pedersen S. (1997) Characterization of the vasodilatatory response to serotonin in human umbilical arteries perfused in vitro. The influence of the endothelium. Early Hum Dev 47:185–193 [DOI] [PubMed] [Google Scholar]
  174. Helke CJ, McDonald CH, Phillips ET. (1993) Hypotensive effects of 5-HT1A receptor activation: ventral medullary sites and mechanisms of action in the rat. J Auton Nerv Syst 42:177–188 [DOI] [PubMed] [Google Scholar]
  175. Hertzler EC. (1961) 5-Hydroxytryptamine and transmission in sympathetic ganglia. Br J Pharmacol 17:406–413 [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Hoffmann P, Carlsson S, Skarphedinsson JO, Thorén P. (1990) Role of different serotonergic receptors in the long-lasting blood pressure depression following muscle stimulation in the spontaneously hypertensive rat. Acta Physiol Scand 139:305–310 [DOI] [PubMed] [Google Scholar]
  177. Holzwarth MA, Brownfield MS. (1985) Serotonin coexists with epinephrine in rat adrenal medullary cells. Neuroendocrinology 41:230–236 [DOI] [PubMed] [Google Scholar]
  178. Holzwarth MA, Sawetawan C, Brownfield MS. (1984) Serotonin-immunoreactivity in the adrenal medulla: distribution and response to pharmacological manipulation. Brain Res Bull 13:299–308 [DOI] [PubMed] [Google Scholar]
  179. Homberg JR, Olivier JD, Smits BM, Mul JD, Mudde J, Verheul M, Nieuwenhuizen OF, Cools AR, Ronken E, Cremers T, et al. (2007) Characterization of the serotonin transporter knockout rat: a selective change in the functioning of the serotonergic system. Neuroscience 146:1662–1676 [DOI] [PubMed] [Google Scholar]
  180. Hoyer D, Clarke DE, Fozard JR, Hartig PR, Martin GR, Mylecharane EJ, Saxena PR, Humphrey PP. (1994) International Union of Pharmacology classification of receptors for 5-hydroxytryptamine (Serotonin). Pharmacol Rev 46:157–203 [PubMed] [Google Scholar]
  181. Hoyer D, Hannon JP, Martin GR. (2002) Molecular, pharmacological and functional diversity of 5-HT receptors. Pharmacol Biochem Behav 71:533–554 [DOI] [PubMed] [Google Scholar]
  182. Huang XP, Setola V, Yadav PN, Allen JA, Rogan SC, Hanson BJ, Revankar C, Robers M, Doucette C, Roth BL. (2009) Parallel functional activity profiling reveals valvulopathogens are potent 5-hydroxytryptamine(2B) receptor agonists: implications for drug safety assessment. Mol Pharmacol 76:710–722 [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Hughes IE, Radwan S. (1979) The relative toxicity of amitriptyline, imipramine, maprotiline and mianserin in rabbits in vivo. Br J Pharmacol 65:331–338 [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Hutri-Kähönen N, Kähönen M, Wu X, Sand J, Nordback I, Taurio J, Pörsti I. (1999) Control of vascular tone in isolated mesenteric arterial segments from hypertensive patients. Br J Pharmacol 127:1735–1743 [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Huzoor-Akbar, Chen NY, Fossen DV, Wallace D. (1989) Increased vascular contractile sensitivity to serotonin in spontaneously hypertensive rats is linked with increased turnover of phosphoinositide. Life Sci 45:577–583 [DOI] [PubMed] [Google Scholar]
  186. Ikeda K, Tojo K, Otsubo C, Udagawa T, Kumazawa K, Ishikawa M, Tokudome G, Hosoya T, Tajima N, Claycomb WC, et al. (2005) 5-Hydroxytryptamine synthesis in HL-1 cells and neonatal rat cardiocytes. Biochem Biophys Res Commun 328:522–525 [DOI] [PubMed] [Google Scholar]
  187. Ishida T, Kawashima S, Hirata Ki, Sakoda T, Shimokawa Y, Miwa Y, Inoue N, Ueyama T, Shiomi M, Akita H, et al. (2001) Serotonin-induced hypercontraction through 5-hydroxytryptamine 1B receptors in atherosclerotic rabbit coronary arteries. Circulation 103:1289–1295 [DOI] [PubMed] [Google Scholar]
  188. Itskovitz HD, Werber JL, Sheridan AM, Brewer TF, Stier CT., Jr (1989) 5-Hydroxytryptophan and carbidopa in spontaneously hypertensive rats. J Hypertens 7:311–315 [PubMed] [Google Scholar]
  189. Jähnichen S, Glusa E, Pertz HH. (2005) Evidence for 5-HT2B and 5-HT7 receptor-mediated relaxation in pulmonary arteries of weaned pigs. Naunyn-Schmiedebergs Arch Pharmacol 371:89–98 [DOI] [PubMed] [Google Scholar]
  190. Janiak P, Lainée P, Grataloup Y, Luyt CE, Bidouard JP, Michel JB, O'Connor SE, Herbert JM. (2002) Serotonin receptor blockade improves distal perfusion after lower limb ischemia in the fatty Zucker rat. Cardiovasc Res 56:293–302 [DOI] [PubMed] [Google Scholar]
  191. Jansen I, Olesen J, Edvinsson L. (1993) 5-Hydroxytryptamine receptor characterization of human cerebral, middle meningeal and temporal arteries: regional differences. Acta Physiol Scand 147:141–150 [DOI] [PubMed] [Google Scholar]
  192. Janssen BJ, van Essen H, Struyker Boudier HA, Smits JF. (1989) Hemodynamic effects of activation of renal and mesenteric sensory nerves in rats. Am J Physiol 257:R29–R36 [DOI] [PubMed] [Google Scholar]
  193. Jelen I, Fananapazir L, Crawford TB. (1979) The possible relation between late pregnancy hypertension and 5-hydroxytryptamine levels in maternal blood. Br J Obstet Gynaecol 86:468–471 [DOI] [PubMed] [Google Scholar]
  194. Jochem J, Zak A, Rybczyk R, Irman-Florjanc T. (2009) Interactions between the serotonergic and histaminergic systems in the central cardiovascular regulation in haemorrhage-shocked rats: involvement of 5-HT(1A) receptors. Inflamm Res 58 (Suppl 1):38–40 [DOI] [PubMed] [Google Scholar]
  195. Johansson K, Sundström J, Neovius K, Rössner S, Neovius M. (2010) Long-term changes in blood pressure following orlistat and sibutramine treatment: a meta-analysis. Obes Rev 11:777–791 [DOI] [PubMed] [Google Scholar]
  196. Jones JF, Martin GR, Ramage AG. (1995) Evidence that 5-HT1D receptors mediate inhibition of sympathetic ganglionic transmission in anaesthetized cats. Br J Pharmacol 116:1715–1717 [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Jordan D. (2005) Vagal control of the heart: central serotonergic (5-HT) mechanisms. Exp Physiol 90:175–181 [DOI] [PubMed] [Google Scholar]
  198. Kamal LA, Le Quan-Bui KH, Meyer P. (1984) Decreased uptake of 3H-serotonin and endogenous content of serotonin in blood platelets in hypertensive patients. Hypertension 6:568–573 [DOI] [PubMed] [Google Scholar]
  199. Karhula T, Soinila S, Lakomy M, Majewski M, Kaleczyk J, Häppölä O. (1995) 5-Hydroxytryptamine-immunoreactive nerve fibers in the rat and porcine vertebral sympathetic ganglia: effect of precursor loading and relation to catecholaminergic neurons. Neurosci Lett 194:85–88 [DOI] [PubMed] [Google Scholar]
  200. Kato H, Noguchi Y, Takagi K. (1974) Comparison of cardiovascular toxicities induced by dimetacrine, imipramine and amitriptyline in isolated guinea pig atria and anesthetized dogs. Jpn J Pharmacol 24:885–891 [DOI] [PubMed] [Google Scholar]
  201. Kaumann AJ, Levy FO. (2006) 5-Hydroxytryptamine receptors in the human cardiovascular system. Pharmacol Ther 111:674–706 [DOI] [PubMed] [Google Scholar]
  202. Kawasaki H, Takasaki K. (1984) Vasoconstrictor response induced by 5-hydroxytryptamine released from vascular adrenergic nerves by periarterial nerve stimulation. J Pharmacol Exp Ther 229:816–822 [PubMed] [Google Scholar]
  203. Kellett DO, Stanford SC, Machado BH, Jordan D, Ramage AG. (2005) Effect of 5-HT depletion on cardiovascular vagal reflex sensitivity in awake and anesthetized rats. Brain Res 1054:61–72 [DOI] [PubMed] [Google Scholar]
  204. Kema IP, Meijer WG, Meiborg G, Ooms B, Willemse PH, de Vries EG. (2001) Profiling of tryptophan-related plasma indoles in patients with carcinoid tumors by automated, on-line, solid-phase extraction and HPLC with fluorescence detection. Clin Chem 47:1811–1820 [PubMed] [Google Scholar]
  205. Kemme MJ, Burggraaf J, Schoemaker RC, Cohen AF, Blauw GJ. (2000) No evidence for functional involvement of 5-HT2B receptors in serotonin-induced vasodilatation in the human forearm. J Cardiovasc Pharmacol 36:699–703 [DOI] [PubMed] [Google Scholar]
  206. Kerecsen L, Bunag RD. (1989) Selective pressor enhancement by monoamine oxidase inhibitors in conscious rats. J Pharmacol Exp Ther 251:645–649 [PubMed] [Google Scholar]
  207. Key BJ, Wigfield CC. (1992) Changes in the tail surface temperature of the rat following injection of 5-hydroxytryptamine into the ventrolateral medulla. Neuropharmacology 31:717–723 [DOI] [PubMed] [Google Scholar]
  208. Kim SH, Lee YM, Jee SH, Nam CM. (2003) Effect of sibutramine on weight loss and blood pressure: a meta-analysis of controlled trials. Obes Res 11:1116–1123 [DOI] [PubMed] [Google Scholar]
  209. Knowles ID, Ramage AG. (1999) Evidence for a role for central 5-HT2B as well as 5-HT2A receptors in cardiovascular regulation in anaesthetized rats. Br J Pharmacol 128:530–542 [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Koch CA, Lasho TL, Tefferi A. (2004) Platelet-rich plasma serotonin levels in chronic myeloproliferative disorders: evaluation of diagnostic use and comparison with the neutrophil PRV-1 assay. Br J Haematol 127:34–39 [DOI] [PubMed] [Google Scholar]
  211. Krobert KA, Bach T, Syversveen T, Kvingedal AM, Levy FO. (2001) The cloned human 5-HT7 receptor splice variants: a comparative characterization of their pharmacology, function and distribution. Naunyn Schmiedebergs Arch Pharmacol 363:620–632 [DOI] [PubMed] [Google Scholar]
  212. Krygicz D, Azzadin A, Pawlak R, Małyszko JS, Pawlak D, Myśliwiec M, Buczko W. (1996) Cyclosporine A affects serotonergic mechanisms in uremic rats. Pol J Pharmacol 48:351–354 [PubMed] [Google Scholar]
  213. Kuhn DM. (1999) Tryptophan hydroxylase regulation. Drug-induced modifications that alter serotonin neuronal function. Adv Exp Med Biol 467:19–27 [DOI] [PubMed] [Google Scholar]
  214. Kuhn DM, Wolf WA, Lovenberg W. (1980) Review of the role of the central serotonergic neuronal system in blood pressure regulation. Hypertension 2:243–255 [DOI] [PubMed] [Google Scholar]
  215. Laguzzi R. (2003) Serotonin2 receptors in the nucleus tractus solitarius: characterization and role in the baroreceptor reflex arc. Cell Mol Neurobiol 23:709–726 [DOI] [PubMed] [Google Scholar]
  216. Lai FM, Tanikella T, Cervoni P. (1991) Characterization of serotonin receptors in isolated rat intramyocardial coronary artery. J Pharmacol Exp Ther 256:164–168 [PubMed] [Google Scholar]
  217. Lamping KG, Faraci FM. (2001) Role of sex differences and effects of endothelial NO synthase deficiency in responses of carotid arteries to serotonin. Arterioscler Thromb Vasc Biol 21:523–528 [DOI] [PubMed] [Google Scholar]
  218. Lazartigues E, Brefel-Courbon C, Bagheri H, Costes S, Gharib C, Tran MA, Senard JM, Montastruc JL. (2000) Fluoxetine-induced pressor response in freely moving rats: a role for vasopressin and sympathetic tone. Fundam Clin Pharmacol 14:443–451 [DOI] [PubMed] [Google Scholar]
  219. Le Quan Sang KH, Montenay-Garestier T, Devynck MA. (1991) Alterations of platelet membrane microviscosity in essential hypertension. Clin Sci 80:205–211 [DOI] [PubMed] [Google Scholar]
  220. Leff P, Martin GR, Morse JM. (1987) Differential classification of vascular smooth muscle and endothelial cell 5-HT receptors by use of tryptamine analogues. Br J Pharmacol 91:321–331 [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Lemberger HF, Mason N, Cohen ML. (1984) 5HT2 receptors in the rat portal vein: desensitization following cumulative serotonin addition. Life Sci 35:71–77 [DOI] [PubMed] [Google Scholar]
  222. Lenglet S, Louiset E, Delarue C, Vaudry H, Contesse V. (2002) Activation of 5-HT(7) receptor in rat glomerulus cells is associated with an increase in adenylyl cyclase activity and calcium influx through T-type calcium channels. Endocrinology 143:1748–1760 [DOI] [PubMed] [Google Scholar]
  223. Lewis DI, Coote JH. (1990) The influence of 5-hydroxytryptamine agonists and antagonists on identified sympathetic preganglionic neurones in the rat, in vivo. Br J Pharmacol 99:667–672 [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Li Kam Wa TC, Freestone S, Samson RR, Johnston NR, Lee MR. (1994) The antinatriuretic action of gamma-L-glutamyl-5-hydroxy-L-tryptophan is dependent on its decarboxylation to 5-hydroxytryptamine in normal man. Br J Clin Pharmacol 38:265–269 [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Lindbom LO, Malatray J, Hall H, Ogren SO. (1982) Cardiovascular and anticholinergic effects of zimeldine. Prog Prog Neuropsychopharmacol Biol Psychiatry 6:403–406 [DOI] [PubMed] [Google Scholar]
  226. Linden A, Desmecht D, Amory H, Lekeux P. (1999) Cardiovascular response to intravenous administration of 5-hydroxytryptamine after type-2 receptor blockade, by metrenperone, in healthy calves. Vet J 157:31–37 [DOI] [PubMed] [Google Scholar]
  227. Linder AE, Beggs KM, Burnett RJ, Watts SW. (2009) Body distribution of infused serotonin in rats. Clin Exp Pharmacol Physiol 36:599–601 [DOI] [PubMed] [Google Scholar]
  228. Linder AE, Diaz J, Ni W, Szasz T, Burnett R, Watts SW. (2008a) Vascular reactivity, 5-HT uptake, and blood pressure in the serotonin transporter knockout rat. Am J Physiol Heart Circ Physiol 294:H1745–H1752 [DOI] [PubMed] [Google Scholar]
  229. Linder AE, Gaskell GL, Szasz T, Thompson JM, Watts SW. (2010) Serotonin receptors in rat jugular vein: presence and involvement in the contraction. J Pharmacol Exp Ther 334:116–123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Linder AE, Ni W, Szasz T, Burnett R, Diaz J, Geddes TJ, Kuhn DM, Watts SW. (2008b) A serotonergic system in veins: serotonin transporter-independent uptake. J Pharmacol Exp Ther 325:714–722 [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Liu GS, Huang DX, Yang XS, Wang SR, Tang XL, Xing JL, Peng SY, Tong YZ. (1991) Ketanserin in the treatment of hypertension: a multicentre dose-finding study in China. Clin Exp Hypertens A 13:541–555 [PubMed] [Google Scholar]
  232. Louiset E, Cartier D, Contesse V, Duparc C, Lihrmann I, Young J, Bertherat J, Reznik Y, Kuhn JM, Laquerrière A, et al. (2004) Paradoxical inhibitory effect of serotonin on cortisol production from adrenocortical lesions causing Cushing's syndrome. Endocr Res 30:951–954 [DOI] [PubMed] [Google Scholar]
  233. MacLean MR, Dempsie Y. (2009) Serotonin and pulmonary hypertension–from bench to bedside? Curr Opin Pharmacol 9:281–286 [DOI] [PubMed] [Google Scholar]
  234. MacLennan SJ, Bolofo ML, Martin GR. (1993) Amplifying interactions between spasmogens in vascular smooth muscle. Biochem Soc Trans 21:1145–1150 [DOI] [PubMed] [Google Scholar]
  235. Madden CJ, Morrison SF. (2006) Serotonin potentiates sympathetic responses evoked by spinal NMDA. J Physiol 577:525–537 [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Manivet P, Mouillet-Richard S, Callebert J, Nebigil CG, Maroteaux L, Hosoda S, Kellermann O, Launay JM. (2000) PDZ-dependent activation of nitric-oxide synthases by the serotonin 2B receptor. J Biol Chem 275:9324–9331 [DOI] [PubMed] [Google Scholar]
  237. Mankad PS, Chester AH, Yacoub MH. (1991) 5-Hydroxytryptamine mediates endothelium dependent coronary vasodilatation in the isolated rat heart by the release of nitric oxide. Cardiovasc Res 25:244–248 [DOI] [PubMed] [Google Scholar]
  238. Martin GR. (1994) Vascular receptors for 5-hydroxytryptamine: distribution, function and classification. Pharmacol Ther 62:283–324 [DOI] [PubMed] [Google Scholar]
  239. Martinez AA, Lokhandwala MF. (1980) Evidence for a presynaptic inhibitory action of 5-hydroxytryptamine on sympathetic neurotransmission to the myocardium. Eur J Pharmacol 63:303–311 [DOI] [PubMed] [Google Scholar]
  240. Marwood JF, Huston V, Wall KT. (1985) Some cardiovascular effects of monoamine oxidase inhibitors in unanaesthetized rats. Clin Exp Pharmacol Physiol 12:161–168 [DOI] [PubMed] [Google Scholar]
  241. Marwood JF, Stokes GS. (1984) Serotonin (5HT) and its antagonists: involvement in the cardiovascular system. Clin Exp Pharmacol Physiol 11:439–455 [DOI] [PubMed] [Google Scholar]
  242. Mason PJ, Morris VA, Balcezak TJ. (2000) Serotonin syndrome. Presentation of 2 cases and review of the literature. Medicine 79:201–209 [DOI] [PubMed] [Google Scholar]
  243. Masu K, Saino T, Kuroda T, Matsuura M, Russa AD, Ishikita N, Satoh Y. (2008) Regional differences in 5-HT receptors in cerebral and testicular arterioles of the rat as revealed by Ca2+ imaging of real-time confocal microscopy: variances by artery size and organ specificity. Arch Histol Cytol 71:291–302 [DOI] [PubMed] [Google Scholar]
  244. Maurer-Spurej E. (2005) Serotonin reuptake inhibitors and cardiovascular diseases: a platelet connection. Cell Mol Life Sci 62:159–170 [DOI] [PubMed] [Google Scholar]
  245. Mbaya P, Alam F, Ashim S, Bennett D. (2007) Cardiovascular effects of high dose venlafaxine XL in patients with major depressive disorder. Hum Psychopharmacol 22:129–133 [DOI] [PubMed] [Google Scholar]
  246. McCall RB, Clement ME. (1994) Role of serotonin1A and serotonin2 receptors in the central regulation of the cardiovascular system. Pharmacol Rev 46:231–243 [PubMed] [Google Scholar]
  247. McGregor DD, Smirk FH. (1970) Vascular responses to 5-hydroxytryptamine in genetic and renal hypertensive rats. Am J Physiol 219:687–690 [DOI] [PubMed] [Google Scholar]
  248. McKune CM, Watts SW. (2001) Characterization of the serotonin receptor mediating contraction in the mouse thoracic aorta and signal pathway coupling. J Pharmacol Exp Ther 297:88–95 [PubMed] [Google Scholar]
  249. McLennan PL, Taylor DA. (1984) Antagonism by ketanserin of 5-HT-induced vasoconstriction unmasks a 5-HT-induced vasodilation. Eur J Pharmacol 104:313–318 [DOI] [PubMed] [Google Scholar]
  250. McMahon FG, Fujioka K, Singh BN, Mendel CM, Rowe E, Rolston K, Johnson F, Mooradian AD. (2000) Efficacy and safety of sibutramine in obese white and African American patients with hypertension: a 1-year, double-blind, placebo-controlled, multicenter trial. Arch Intern Med 160:2185–2191 [DOI] [PubMed] [Google Scholar]
  251. Medow MS, Stewart JM, Sanyal S, Mumtaz A, Sica D, Frishman WH. (2008) Pathophysiology, diagnosis, and treatment of orthostatic hypotension and vasovagal syncope. Cardiol Rev 16:4–20 [DOI] [PubMed] [Google Scholar]
  252. Meehan AG, Kreulen DL. (1991) Electrophysiological studies on the interaction of 5-hydroxytryptamine with sympathetic transmission in the guinea pig inferior mesenteric artery and ganglion. J Pharmacol Exp Ther 256:82–87 [PubMed] [Google Scholar]
  253. Merahi N, Laguzzi R. (1995) Cardiovascular effects of 5HT2 and 5HT3 receptor stimulation in the nucleus tractus solitarius of spontaneously hypertensive rats. Brain Res 669:130–134 [DOI] [PubMed] [Google Scholar]
  254. Mercado CP, Ziu E, Kilic F. (2011) Communication between 5-HT and small GTPases. Curr Opin Pharmacol 11:23–28 [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Meyer T, Brinck U. (1999) Differential distribution of serotonin and tryptophan hydroxylase in the human gastrointestinal tract. Digestion 60:63–68 [DOI] [PubMed] [Google Scholar]
  256. Miller M, Hasson R, Morgane PJ, Resnick O. (1980) Adrenalectomy: its effects on systemic tryptophan metabolism in normal and protein malnourished rats. Brain Res Bull 5:451–459 [DOI] [PubMed] [Google Scholar]
  257. Miller RC, Cornish EJ, Goldie RG. (1984) Responses of sheep and kitten isolated coronary arteries to some vasoactive agents. Pharmacol Res Commun 16:667–677 [DOI] [PubMed] [Google Scholar]
  258. Miranda FJ, Torregrosa G, Salom JB, Alabadí JA, Jover T, Barberá MD, Alborch E. (1995) Characterization of 5-hydroxytryptamine receptors in goat cerebral arteries. Gen Pharmacol 26:1267–1272 [DOI] [PubMed] [Google Scholar]
  259. Molderings GJ, Werner K, Likungu J, Göthert M. (1990) Inhibition of noradrenaline release from the sympathetic nerves of the human saphenous vein via presynaptic 5-HT receptors similar to the 5-HT 1D subtype. Naunyn Schmiedebergs Arch Pharmacol 342:371–377 [DOI] [PubMed] [Google Scholar]
  260. Monaghan PJ, Brown HA, Houghton LA, Keevil BG. (2009) Measurement of serotonin in platelet depleted plasma by liquid chromatography tandem mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 877:2163–2167 [DOI] [PubMed] [Google Scholar]
  261. Monassier L, Laplante MA, Ayadi T, Doly S, Maroteaux L. (2010) Contribution of gene-modified mice and rats to our understanding of the cardiovascular pharmacology of serotonin. Pharmacol Ther 128:559–567 [DOI] [PubMed] [Google Scholar]
  262. Montes R, Johnson AK. (1990) Efferent mechanisms mediating renal sodium and water excretion induced by centrally administered serotonin. Am J Physiol 259:R1267–R1273 [DOI] [PubMed] [Google Scholar]
  263. Morán A, Restrepo B, de Urbina AV, García M, Martín ML, Román LS. (2010) Pharmacological profile of 5-hydroxytryptamine-induced inhibition on the pressor effect elicited by sympathetic stimulation in long-term diabetic pithed rats. Eur J Pharmacol 643:70–77 [DOI] [PubMed] [Google Scholar]
  264. Morecroft I, Dempsie Y, Bader M, Walther DJ, Kotnik K, Loughlin L, Nilsen M, MacLean MR. (2007) Effect of tryptophan hydroxylase 1 deficiency on the development of hypoxia-induced pulmonary hypertension. Hypertension 49:232–236 [DOI] [PubMed] [Google Scholar]
  265. Moreira TS, Takakura AC, Colombari E, Guyenet PG. (2007) Activation of 5-hydroxytryptamine type 3 receptor-expressing C-fiber vagal afferents inhibits retrotrapezoid nucleus chemoreceptors in rats. J Neurophysiol 98:3627–3637 [DOI] [PubMed] [Google Scholar]
  266. Moreno L, Martínez-Cuesta MA, Piqué JM, Bosch J, Esplugues JV. (1996) Anatomical differences in responsiveness to vasoconstrictors in the mesenteric veins from normal and portal hypertensive rats. Naunyn Schmiedebergs Arch Pharmacol 354:474–480 [DOI] [PubMed] [Google Scholar]
  267. Morrison SF. (2004) Central pathways controlling brown adipose tissue thermogenesis. News Physiol Sci 19:67–74 [DOI] [PubMed] [Google Scholar]
  268. Murphy DL, Lipper S, Slater S, Shiling D. (1979) Selectivity of clorgyline and pargyline as inhibitors of monoamine oxidases A and B in vivo in man. Psychopharmacology 62:129–132 [DOI] [PubMed] [Google Scholar]
  269. Mylecharane EJ. (1990) Mechanisms involved in serotonin-induced vasodilatation. Blood Vessels 27:116–126 [DOI] [PubMed] [Google Scholar]
  270. Nagai S, Tsurumaki T, Abe H, Higuchi H. (2007) Functional serotonin and histamine receptor subtypes in porcine ciliary artery in comparison with middle cerebral artery. Eur J Pharmacol 570:159–166 [DOI] [PubMed] [Google Scholar]
  271. Nagatsu I, Ichinose H, Sakai M, Titani K, Suzuki M, Nagatsu T. (1995) Immunocytochemical localization of GTP cyclohydrolase I in the brain, adrenal gland, and liver of mice. J Neural Transm Gen Sect 102:175–188 [DOI] [PubMed] [Google Scholar]
  272. Nakatani Y, Sato-Suzuki I, Tsujino N, Nakasato A, Seki Y, Fumoto M, Arita H. (2008) Augmented brain 5-HT crosses the blood-brain barrier through the 5-HT transporter in rat. Eur J Neurosci 27:2466–2472 [DOI] [PubMed] [Google Scholar]
  273. Nalivaiko E, Sgoifo A. (2009) Central 5-HT receptors in cardiovascular control during stress. Neurosci Biobehav Rev 33:95–106 [DOI] [PubMed] [Google Scholar]
  274. Nebigil CG, Etienne N, Messaddeq N, Maroteaux L. (2003a) Serotonin is a novel survival factor of cardiomyocytes: mitochondria as a target of 5-HT2B receptor signaling. FASEB J 17:1373–1375 [DOI] [PubMed] [Google Scholar]
  275. Nebigil CG, Jaffré F, Messaddeq N, Hickel P, Monassier L, Launay JM, Maroteaux L. (2003b) Overexpression of the serotonin 5-HT2B receptor in heart leads to abnormal mitochondrial function and cardiac hypertrophy. Circulation 107:3223–3229 [DOI] [PubMed] [Google Scholar]
  276. Nebigil CG, Maroteaux L. (2001) A novel role for serotonin in heart. Trends Cardiovasc Med 11:329–335 [DOI] [PubMed] [Google Scholar]
  277. Nelson M, Coghlan JP, Denton DA, Tresham JJ, Whitworth JA, Scoggins BA. (1987) Ritanserin and serotonergic mechanisms in blood pressure and fluid regulation in sheep. Clin Exp Pharmacol Physiol 14:555–563 [DOI] [PubMed] [Google Scholar]
  278. Newberry NR, Watkins CJ, Volenec A, Flanigan TP. (1996) 5-HT2B receptor mRNA in guinea pig superior cervical ganglion. Neuroreport 7:2909–2911 [DOI] [PubMed] [Google Scholar]
  279. Ni W, Geddes TJ, Priestley JR, Szasz T, Kuhn DM, Watts SW. (2008) The existence of a local 5-hydroxytryptaminergic system in peripheral arteries. Br J Pharmacol 154:663–674 [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Ni W, Thompson JM, Northcott CA, Lookingland K, Watts SW. (2004) The serotonin transporter is present and functional in peripheral arterial smooth muscle. J Cardiovasc Pharmacol 43:770–781 [DOI] [PubMed] [Google Scholar]
  281. Nichols CD. (2009) Serotonin 5-HT2A receptor function as a contributing factor to both neuropsychiatric and cardiovascular diseases. Cardiovasc Psychiatry Neurol 2009:475108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  282. Nigmatullina RR, Kirillova VV, Jourjikiya RK, Mukhamedyarov MA, Kudrin VS, Klodt PM, Palotás A. (2009) Disrupted serotonergic and sympathoadrenal systems in patients with chronic heart failure may serve as new therapeutic targets and novel biomarkers to assess severity, progression and response to treatment. Cardiology 113:277–286 [DOI] [PubMed] [Google Scholar]
  283. Nilsson T, Longmore J, Shaw D, Olesen IJ, Edvinsson L. (1999) Contractile 5-HT1B receptors in human cerebral arteries: pharmacological characterization and localization with immunocytochemistry. Br J Pharmacol 128:1133–1140 [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Nishimura M, Sato K, Shimada S, Tohyama M. (1999) Expression of norepinephrine and serotonin transporter mRNAs in the rat superior cervical ganglion. Brain Res Mol Brain Res 67:82–86 [DOI] [PubMed] [Google Scholar]
  285. Nishimura Y. (1996) Characterization of 5-hydroxytryptamine receptors mediating contractions in basilar arteries from stroke-prone spontaneously hypertensive rats. Br J Pharmacol 117:1325–1333 [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Nishimura Y, Suzuki A. (1995) Enhanced contractile responses mediated by different 5-HT receptor subtypes in basilar arteries, superior mesenteric arteries and thoracic aortas from stroke-prone spontaneously hypertensive rats. Clin Exp Pharmacol Physiol Suppl 22:S99–S101 [DOI] [PubMed] [Google Scholar]
  287. Nolen WA, Haffmans PM, Bouvy PF, Duivenvoorden HJ. (1993) Monoamine oxidase inhibitors in resistant major depression. A double-blind comparison of brofaromine and tranylcypromine in patients resistant to tricyclic antidepressants. J Affect Disord 28:189–197 [DOI] [PubMed] [Google Scholar]
  288. Nosjean A, Compoint C, Buisseret-Delmas C, Orer HS, Merahi N, Puizillout JJ, Laguzzi R. (1990) Serotonergic projections from the nodose ganglia to the nucleus tractus solitarius: an immunohistochemical and double labeling study in the rat. Neurosci Lett 114:22–26 [DOI] [PubMed] [Google Scholar]
  289. Nosjean A, Franc B, Laguzzi R. (1995) Increased sympathetic nerve discharge without alteration in the sympathetic baroreflex response by serotonin3 receptor stimulation in the nucleus tractus solitarius of the rat. Neurosci Lett 186:41–44 [DOI] [PubMed] [Google Scholar]
  290. Nosjean A, Guyenet PG. (1991) Role of ventrolateral medulla in sympatholytic effect of 8-OHDPAT in rats. Am J Physiol 260:R600–R609 [DOI] [PubMed] [Google Scholar]
  291. Nyborg NC, Mikkelsen EO. (1985) In vitro studies on responses to noradrenaline, serotonin, and potassium of intramyocardial and mesenteric resistance vessels from Wistar rats. J Cardiovasc Pharmacol 7:417–423 [DOI] [PubMed] [Google Scholar]
  292. Obi T, Kabeyama A, Nishio A. (1994) Equine coronary artery responds to 5-hydroxytryptamine with relaxation in vitro. J Vet Pharmacol Ther 17:218–225 [DOI] [PubMed] [Google Scholar]
  293. Ogura C, Kishimoto A, Mizukawa R, Hazama H, Honma H, Kawahara K. (1983) Age differences in effects on blood pressure, flicker fusion frequency, salivation and pharmacokinetics of single oral doses of dothiepin and amitriptyline. Eur J Clin Pharmacol 25:811–814 [DOI] [PubMed] [Google Scholar]
  294. Ohtsuki S. (2004) New aspects of the blood-brain barrier transporters; its physiological roles in the central nervous system. Biol Pharm Bull 27:1489–1496 [DOI] [PubMed] [Google Scholar]
  295. Ootsuka Y, Blessing WW. (2005) Activation of slowly conducting medullary raphe-spinal neurons, including serotonergic neurons, increases cutaneous sympathetic vasomotor discharge in rabbit. Am J Physiol Regul Integr Comp Physiol 288:R909–R918 [DOI] [PubMed] [Google Scholar]
  296. Ootsuka Y, Blessing WW. (2006) Activation of 5-HT1A receptors in rostral medullary raphe inhibits cutaneous vasoconstriction elicited by cold exposure in rabbits. Brain Res 1073–1074:252–261 [DOI] [PubMed] [Google Scholar]
  297. Orer HS, Merahi N, Nosjean A, Fattaccini CM, Laguzzi R. (1991) Cardiovascular effects of the local injection of 5,7-dihydroxytryptamine into the nodose ganglia and nucleus tractus solitarius in awake freely moving rats. Brain Res 553:123–128 [DOI] [PubMed] [Google Scholar]
  298. Page IH, McCubbin JW. (1953a) The variable arterial pressure response to serotonin in laboratory animals and man. Circ Res 1:354–362 [DOI] [PubMed] [Google Scholar]
  299. Page IH, McCubbin JW. (1953b) Modification of vascular responses to serotonin by drugs. Am J Physiol 174:436–444 [DOI] [PubMed] [Google Scholar]
  300. Pagniez F, Valentin JP, Vieu S, Colpaert FC, John GW. (1998) Pharmacological analysis of the haemodynamic effects of 5-HT1B/D receptor agonists in the normotensive rat. Br J Pharmacol 123:205–214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  301. Päivärinta H, Park DH, Towle AC, Joh TH. (1989) Tryptophan hydroxylase activity and 5-hydroxytryptamine-immunoreactive cells in the superior cervical ganglion of hydrocortisone-treated neonatal rats. Neurosci Res 6:276–281 [DOI] [PubMed] [Google Scholar]
  302. Park CS, Chu CS, Park YS, Hong SK. (1968) Effect of 5-hydroxytryptamine on renal function of the anesthetized dog. Am J Physiol 214:383–388 [DOI] [PubMed] [Google Scholar]
  303. Parsons AA. (1991) 5-HT receptors in human and animal cerebrovasculature. Trends Pharmacol Sci 12:310–315 [DOI] [PubMed] [Google Scholar]
  304. Parsons AA, Raval P, Smith S, Tilford N, King FD, Kaumann AJ, Hunter J. (1998) Effects of the novel high-affinity 5-HT(1B/1D)-receptor ligand frovatriptan in human isolated basilar and coronary arteries. J Cardiovasc Pharmacol 32:220–224 [DOI] [PubMed] [Google Scholar]
  305. Parsons AA, Stutchbury C, Raval P, Kaumann AJ. (1992) Sumatriptan contracts large coronary arteries of beagle dogs through 5-HT1-like receptors. Naunyn Schmiedebergs Arch Pharmacol 346:592–596 [DOI] [PubMed] [Google Scholar]
  306. Parsons AA, Whalley ET, Feniuk W, Connor HE, Humphrey PP. (1989) 5-HT1-like receptors mediate 5-hydroxytryptamine-induced contraction of human isolated basilar artery. Br J Pharmacol 96:434–440 [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Paulmann N, Grohmann M, Voigt JP, Bert B, Vowinckel J, Bader M, Skelin M, Jevsek M, Fink H, Rupnik M, et al. (2009) Intracellular serotonin modulates insulin secretion from pancreatic β-cells by protein serotonylation. PLoS Biol 7:e1000229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Pavone LM, Spina A, Lo Muto R, Santoro D, Mastellone V, Avallone L. (2008) Heart valve cardiomyocytes of mouse embryos express the serotonin transporter SERT. Biochem Biophys Res Commun 377:419–422 [DOI] [PubMed] [Google Scholar]
  309. Pavone LM, Spina A, Rea S, Santoro D, Mastellone V, Lombardi P, Avallone L. (2009) Serotonin transporter gene deficiency is associated with sudden death of newborn mice through activation of TGF-beta1 signalling. J Mol Cell Cardiol 47:691–697 [DOI] [PubMed] [Google Scholar]
  310. PDSP Ki Database; http://pdsp.med.unc.edu/indexR.html
  311. Pedersen OL, Kragh-Sørensen P, Bjerre M, Overø KF, Gram LF. (1982) Citalopram, a selective serotonin reuptake inhibitor: clinical antidepressive and long-term effect–a phase II study. Psychopharmacology 77:199–204 [DOI] [PubMed] [Google Scholar]
  312. Penttilä J, Syvälahti E, Hinkka S, Kuusela T, Scheinin H. (2001) The effects of amitriptyline, citalopram and reboxetine on autonomic nervous system. A randomised placebo-controlled study on healthy volunteers. Psychopharmacology 154:343–349 [DOI] [PubMed] [Google Scholar]
  313. Perahia DG, Quail D, Gandhi P, Walker DJ, Peveler RC. (2008) A randomized, controlled trial of duloxetine alone vs. duloxetine plus a telephone intervention in the treatment of depression. J Affect Disord 108:33–41 [DOI] [PubMed] [Google Scholar]
  314. Persson B, Gradin K, Pettersson A, Hedner T. (1988) Antihypertensive effects of ketanserin and ritanserin in the spontaneously hypertensive rat. J Cardiovasc Pharmacol 11(S1):S22–S24 [PubMed] [Google Scholar]
  315. Phillips CA, Mylecharane EJ, Shaw J. (1985) Mechanisms involved in the vasodilator action of 5-hydroxytryptamine in the dog femoral arterial circulation in vivo. Eur J Pharmacol 113:325–334 [DOI] [PubMed] [Google Scholar]
  316. Pickering AE, Spanswick D, Logan SD. (1994) 5-Hydoxytryptamine evokes depolarizations and membrane potential oscillations in rat sympathetic preganglionic neurones. J Physiol 480:109–121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  317. Pierce PA, Xie GX, Levine JD, Peroutka SJ. (1996) 5-Hydroxytryptamine receptor subtype messenger RNAs in rat peripheral sensory and sympathetic ganglia: a polymerase chain reaction study. Neuroscience 70:553–559 [DOI] [PubMed] [Google Scholar]
  318. Pozdzik M, Zajac D, Zasada I, Czarnocki Z, Matysiak Z, Mazzatenta A, Pokorski M. (2010) Absence of bioactivity of lipid derivatives of serotonin. Eur J Med Res 15:128–134 [DOI] [PMC free article] [PubMed] [Google Scholar]
  319. Radenkoviä M, Stojanoviä M, Topaloviä M. (2010) Contribution of thromboxane A2 in rat common carotid artery response to serotonin. Sci Pharm 78:435–443 [DOI] [PMC free article] [PubMed] [Google Scholar]
  320. Rajapakse S, Abeynaike L, Wickramarathne T. (2010) Venlafaxine-associated serotonin syndrome causing severe rhabdomyolysis and acute renal failure in a patient with idiopathic Parkinson disease. J Clin Psychopharmacol 30:620–622 [DOI] [PubMed] [Google Scholar]
  321. Ramage AG. (2001) Central cardiovascular regulation and 5-hydroxytryptamine receptors. Brain Res Bull 56:425–439 [DOI] [PubMed] [Google Scholar]
  322. Ramage AG, Daly MB. (1998) The central action of the 5-HT2 receptor agonist 1-(2,5-dimethoxy-4-iodophenyl)-2-aminopropane (DOI) on cardiac inotropy and vascular resistance in the anaesthetized cat. Br J Pharmacol 125:1172–1179 [DOI] [PMC free article] [PubMed] [Google Scholar]
  323. Ramage AG, Fozard JR. (1987) Evidence that the putative 5-HT1A receptor agonists, 8-OH-DPAT and ipsapirone, have a central hypotensive action that differs from that of clonidine in anaesthetised cats. Eur J Pharmacol 138:179–191 [DOI] [PubMed] [Google Scholar]
  324. Ramage AG, Villalón CM. (2008) 5-Hydroxytryptamine and cardiovascular regulation. Trends Pharmacol Sci 29:472–481 [DOI] [PubMed] [Google Scholar]
  325. Rapport MM, Green AA, Page IH. (1948) Serum vasoconstrictor, serotonin; isolation and characterization. J Biol Chem 176:1243–1251 [PubMed] [Google Scholar]
  326. Razzaque Z, Pickard JD, Ma QP, Shaw D, Morrison K, Wang T, Longmore J. (2002) 5-HT1B-receptors and vascular reactivity in human isolated blood vessels: assessment of the potential craniovascular selectivity of sumatriptan. Br J Clin Pharmacol 53:266–274 [DOI] [PMC free article] [PubMed] [Google Scholar]
  327. Roon KI, Maassen Van Den Brink A, Ferrari MD, Saxena PR. (1999) Bovine isolated middle cerebral artery contractions to antimigraine drugs. Naunyn Schmiedebergs Arch Pharmacol 360:591–596 [DOI] [PubMed] [Google Scholar]
  328. Rosón MI, Maquieira MK, de la Riva IJ. (1990) Contrasting effects of norepinephrine and 5-hydroxytryptamine on contractility of abdominal aorta of two kidney-two clip hypertensive rats. Effects of inhibitors of arachidonic acid metabolic enzymes. Clin Exp Hypertens A 12:285–306 [DOI] [PubMed] [Google Scholar]
  329. Ross RJ, Zavadil AP, 3rd, Calil HM, Linnoila M, Kitanaka I, Blombery P, Kopin IJ, Potter WZ. (1983) Effects of desmethylimipramine on plasma norepinephrine, pulse, and blood pressure. Clin Pharmacol Ther 33:429–437 [DOI] [PubMed] [Google Scholar]
  330. Roth BL. (2007) Drugs and valvular heart disease. N Engl J Med 356:6–9 [DOI] [PubMed] [Google Scholar]
  331. Roux F, Couraud PO. (2005) Rat brain endothelial cell lines for the study of blood-brain barrier permeability and transport functions. Cell Mol Neurobiol 25:41–58 [DOI] [PubMed] [Google Scholar]
  332. Russell A, Banes A, Berlin H, Fink GD, Watts SW. (2002) 5-Hydroxytryptamine(2B) receptor function is enhanced in the N(omega)-nitro-L-arginine hypertensive rat. J Pharmacol Exp Ther 303:179–187 [DOI] [PubMed] [Google Scholar]
  333. Saletu B, Grünberger J, Linzmayer L. (1986) On central effects of serotonin re-uptake inhibitors: quantitative EEG and psychometric studies with sertraline and zimelidine. J Neural Transm 67:241–266 [DOI] [PubMed] [Google Scholar]
  334. Salter M, Hazelwood R, Pogson CI, Iyer R, Madge DJ. (1995) The effects of a novel and selective inhibitor of tryptophan 2,3-dioxygenase on tryptophan and serotonin metabolism in the rat. Biochem Pharmacol 49:1435–1442 [DOI] [PubMed] [Google Scholar]
  335. Sari Y, Zhou FC. (2003) Serotonin and its transporter on proliferation of fetal heart cells. Int J Dev Neurosci 21:417–424 [DOI] [PubMed] [Google Scholar]
  336. Savelieva KV, Zhao S, Pogorelov VM, Rajan I, Yang Q, Cullinan E, Lanthorn TH. (2008) Genetic disruption of both tryptophan hydroxylase genes dramatically reduces serotonin and affects behavior in models sensitive to antidepressants. PLoS ONE 3:e3301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  337. Saxena PR, Verdouw PD. (1982) Redistribution by 5-hydroxytryptamine of carotid arterial blood at the expense of arteriovenous anastomotic blood flow. J Physiol 332:501–520 [DOI] [PMC free article] [PubMed] [Google Scholar]
  338. Saxena PR, Villalón CM. (1991) 5-Hydroxytryptamine: a chameleon in the heart. Trends Pharmacol Sci 12:223–227 [DOI] [PubMed] [Google Scholar]
  339. Saydoff JA, Rittenhouse PA, Carnes M, Armstrong J, Van De Kar LD, Brownfield MS. (1996) Neuroendocrine and cardiovascular effects of serotonin: selective role of brain angiotensin on vasopressin. Am J Physiol 270:E513–E521 [DOI] [PubMed] [Google Scholar]
  340. Schmuck K, Ullmer C, Kalkman HO, Probst A, Lubbert H. (1996) Activation of meningeal 5-HT2B receptors: an early step in the generation of migraine headache? Eur J Neurosci 8:959–967 [DOI] [PubMed] [Google Scholar]
  341. Schöning C, Flieger M, Pertz HH. (2001) Complex interaction of ergovaline with 5-HT2A, 5-HT1B/1D, and alpha1 receptors in isolated arteries of rat and guinea pig. J Anim Sci 79:2202–2209 [DOI] [PubMed] [Google Scholar]
  342. Scrogin KE. (2003) 5-HT1A receptor agonist 8-OH-DPAT acts in the hindbrain to reverse the sympatholytic response to severe hemorrhage. Am J Physiol Regul Integr Comp Physiol 284:R782–R791 [DOI] [PubMed] [Google Scholar]
  343. Sharma AM, Caterson ID, Coutinho W, Finer N, Van Gaal L, Maggioni AP, Torp-Pedersen C, Bacher HP, Shepherd GM, James WP, et al. (2009) Blood pressure changes associated with sibutramine and weight management - an analysis from the 6-week lead-in period of the sibutramine cardiovascular outcomes trial (SCOUT). Diabetes Obes Metab 11:239–250 [DOI] [PubMed] [Google Scholar]
  344. Sharma HS, Dey PK. (1984) Role of 5-HT on increased permeability of blood-brain barrier under heat stress. Indian J Physiol Pharmacol 28:259–267 [PubMed] [Google Scholar]
  345. Shepheard SL, Jordan D, Ramage AG. (1994) Comparison of the effects of IVth ventricular administration of some tryptamine analogues with those of 8-OH-DPAT on autonomic outflow in the anaesthetized cat. Br J Pharmacol 111:616–624 [DOI] [PMC free article] [PubMed] [Google Scholar]
  346. Shepherd JT, Vanhoutte PM. (1985) Local modulation of adrenergic neurotransmission in blood vessels. J Cardiovasc Pharmacol 7 (Suppl 3):S167–S178 [DOI] [PubMed] [Google Scholar]
  347. Shoji T, Tamaki T, Fukui K, Iwao H, Abe Y. (1989) Renal hemodynamic responses to 5-hydroxytryptamine (5-HT): involvement of the 5-HT receptor subtypes in the canine kidney. Eur J Pharmacol 171:219–228 [DOI] [PubMed] [Google Scholar]
  348. Skop BP, Finkelstein JA, Mareth TR, Magoon MR, Brown TM. (1994) The serotonin syndrome associated with paroxetine, an over-the-counter cold remedy, and vascular disease. Am J Emerg Med 12:642–644 [DOI] [PubMed] [Google Scholar]
  349. Soares-da-Silva P, Vieira-Coelho MA, Pestana M. (1996) Antagonistic actions of renal dopamine and 5-hydroxytryptamine: endogenous 5-hydroxytryptamine, 5-HT1A receptors and antinatriuresis during high sodium intake. Br J Pharmacol 117:1193–1198 [DOI] [PMC free article] [PubMed] [Google Scholar]
  350. Sole MJ, Madapallimattam A, Baines AD. (1986) An active pathway for serotonin synthesis by renal proximal tubules. Kidney Int 29:689–694 [DOI] [PubMed] [Google Scholar]
  351. Sole MJ, Shum A, Van Loon GR. (1979) Serotonin metabolism in the normal and failing hamster heart. Circ Res 45:629–634 [DOI] [PubMed] [Google Scholar]
  352. Sprouse J, Reynolds L, Li X, Braselton J, Schmidt A. (2004) 8-OH-DPAT as a 5-HT7 agonist: phase shifts of the circadian biological clock through increases in cAMP production. Neuropharmacology 46:52–62 [DOI] [PubMed] [Google Scholar]
  353. Stahl SM, Felker A. (2008) Monoamine oxidase inhibitors: a modern guide to an unrequited class of antidepressants. CNS Spectr 13:855–870 [DOI] [PubMed] [Google Scholar]
  354. Stevens H, Knollema S, De Jong G, Korf J, Luiten PG. (1998) Long-term food restriction, deprenyl, and nimodipine treatment on life expectancy and blood pressure of stroke-prone rats. Neurobiol Aging 19:273–276 [DOI] [PubMed] [Google Scholar]
  355. Stier CT, Jr, Itskovitz HD. (1985) Formation of serotonin by rat kidneys in vivo. Proc Soc Exp Biol. Med 180:550–557 [DOI] [PubMed] [Google Scholar]
  356. Stier CT, Jr, Brewer TF, Dick LB, Wynn N, Itskovitz HD. (1986) Formation of biogenic amines by isolated kidneys of spontaneously hypertensive rats. Life Sci 38:7–14 [DOI] [PubMed] [Google Scholar]
  357. Stone TW, Darlington LG. (2002) Endogenous kynurenines as targets for drug discovery and development. Nat Rev Drug Discov 1:609–620 [DOI] [PubMed] [Google Scholar]
  358. Stott DJ, Saniabadi AR, Hosie J, Lowe GD, Ball SG. (1988) The effects of the 5 HT2 antagonist ritanserin on blood pressure and serotonin-induced platelet aggregation in patients with untreated essential hypertension. Eur J Clin Pharmacol 35:123–129 [DOI] [PubMed] [Google Scholar]
  359. Subramanian S, Vollmer RR. (2004) Fenfluramine-induced hypothermia is associated with cutaneous dilation in conscious rats. Pharmacol Biochem Behav 77:351–359 [DOI] [PubMed] [Google Scholar]
  360. Sugimoto Y, Yamada J, Yoshikawa T, Nishikawa F, Noma T, Horisaka K. (1996) The involvement of serotonin in the catecholamine release from the adrenal medulla. Adv Exp Med Biol 398:561–563 [DOI] [PubMed] [Google Scholar]
  361. Sumner MJ. (1991) Characterization of the 5-HT receptor mediating endothelium-dependent relaxation in porcine vena cava. Br J Pharmacol 102:938–942 [DOI] [PMC free article] [PubMed] [Google Scholar]
  362. Sved AF, Van Itallie CM, Fernstrom JD. (1982) Studies on the antihypertensive action of L-tryptophan. J Pharmacol Exp Ther 221:329–333 [PubMed] [Google Scholar]
  363. Takeuchi Y, Sano Y. (1983) Serotonin distribution in the circumventricular organs of the rat. An immunohistochemical study. Anat Embryol 167:311–319 [DOI] [PubMed] [Google Scholar]
  364. Tan T, Watts SW, Davis RP. (2011) Drug delivery: enabling technology for drug discovery and development. iPRECIO micro infusion pump: programmable, refillable, and implantable. Front Pharmacol 2:44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  365. Teng GQ, Nauli SM, Brayden JE, Pearce WJ. (2002) Maturation alters the contribution of potassium channels to resting and 5HT-induced tone in small cerebral arteries of the sheep. Brain Res Dev Brain Res 133:81–91 [DOI] [PubMed] [Google Scholar]
  366. Terrón JA. (1996) The relaxant 5-HT receptor in the dog coronary artery smooth muscle: pharmacological resemblance to the cloned 5-ht7 receptor subtype. Br J Pharmacol 118:1421–1428 [DOI] [PMC free article] [PubMed] [Google Scholar]
  367. Terrón JA. (1997) Role of 5-ht7 receptors in the long-lasting hypotensive response induced by 5-hydroxytryptamine in the rat. Br J Pharmacol 121:563–571 [DOI] [PMC free article] [PubMed] [Google Scholar]
  368. Terrón JA, Falcón-Neri A. (1999) Pharmacological evidence for the 5-HT7 receptor mediating smooth muscle relaxation in canine cerebral arteries. Br J Pharmacol 127:609–616 [DOI] [PMC free article] [PubMed] [Google Scholar]
  369. Terrón JA, Sánchez-Maldonado C, Martínez-García E. (2007) Pharmacological evidence that 5-HT(1B/1D) receptors mediate hypotension in anesthetized rats. Eur J Pharmacol 576:132–135 [DOI] [PubMed] [Google Scholar]
  370. Thase ME. (1998) Effects of venlafaxine on blood pressure: a meta-analysis of original data from 3744 depressed patients. J Clin Psychiatry 59:502–508 [DOI] [PubMed] [Google Scholar]
  371. Thayssen P, Bjerre M, Kragh-Sørensen P, Møller M, Petersen OL, Kristensen CB, Gram LF. (1981) Cardiovascular effect of imipramine and nortriptyline in elderly patients. Psychopharmacology 74:360–364 [DOI] [PubMed] [Google Scholar]
  372. Thompson LP, Webb RC. (1987) Vascular responsiveness to serotonin metabolites in mineralocorticoid hypertension. Hypertension 9:277–281 [DOI] [PubMed] [Google Scholar]
  373. Thor KB, Helke CJ. (1987) Serotonin- and substance P-containing projections to the nucleus tractus solitarii of the rat. J Comp Neurol 265:275–293 [DOI] [PubMed] [Google Scholar]
  374. Tiniakov R, Osei-Owusu P, Scrogin KE. (2007) The 5-hydroxytryptamine1A receptor agonist, (+)-8-hydroxy-2-(di-n-propylamino)-tetralin, increases cardiac output and renal perfusion in rats subjected to hypovolemic shock. J Pharmacol Exp Ther 320:811–818 [DOI] [PubMed] [Google Scholar]
  375. Toda N, Okamura T. (1990) Comparison of the response to 5-carboxamidotryptamine and serotonin in isolated human, monkey and dog coronary arteries. J Pharmacol Exp Ther 253:676–682 [PubMed] [Google Scholar]
  376. Topsakal R, Kalay N, Gunturk EE, Dogan A, Inanc MT, Kaya MG, Ergin A, Yarlioglues M. (2009) The relation between serotonin levels and insufficient blood pressure decrease during night-time in hypertensive patients. Blood Press 18:367–371 [DOI] [PubMed] [Google Scholar]
  377. Tsai ML, Lin MT. (1986) Hypertension and tachycardia produced by inhibition of reuptake of 5-hydroxytryptamine by fluoxetine in the rat. Neuropharmacology 25:799–802 [DOI] [PubMed] [Google Scholar]
  378. Turkka J, Suominen K, Tolonen U, Sotaniemi K, Myllylä VV. (1997) Selegiline diminishes cardiovascular autonomic responses in Parkinson's disease. Neurology 48:662–667 [DOI] [PubMed] [Google Scholar]
  379. Turner EH, Loftis JM, Blackwell AD. (2006) Serotonin a la carte: supplementation with the serotonin precursor 5-hydroxytryptophan. Pharmacol Ther 109:325–338 [DOI] [PubMed] [Google Scholar]
  380. Tytgat J, Maertens C, Daenens P. (1997) Effect of fluoxetine on a neuronal, voltage-dependent potassium channel (Kv1.1). Br J Pharmacol 122:1417–1424 [DOI] [PMC free article] [PubMed] [Google Scholar]
  381. Ueno M. (2009) Mechanisms of the penetration of blood-borne substances into the brain. Curr Neuropharmacol 7:142–149 [DOI] [PMC free article] [PubMed] [Google Scholar]
  382. Umegaki K, Nakamura K, Ikeda M, Inoue Y, Tomita T. (1989) Changes in platelet function due to hypertension: comparison of experimental hypertension with spontaneous hypertension in rats. J Hypertens 7:13–19 [DOI] [PubMed] [Google Scholar]
  383. Valenta B, Singer EA. (1990) Hypotensive effects of 8-hydroxy-2-(di-n-propylamino)tetralin and 5-methylurapidil following stereotaxic microinjection into the ventral medulla of the rat. Br J Pharmacol 99:713–716 [DOI] [PMC free article] [PubMed] [Google Scholar]
  384. Valentin JP, Bonnafous R, John GW. (1996) Influence of the endothelium and nitric oxide on the contractile responses evoked by 5-HT1D receptor agonists in the rabbit isolated saphenous vein. Br J Pharmacol 119:35–42 [DOI] [PMC free article] [PubMed] [Google Scholar]
  385. van den Broek RW, Bhalla P, VanDenBrink AM, de Vries R, Sharma HS, Saxena PR. (2002) Characterization of sumatriptan-induced contractions in human isolated blood vessels using selective 5-HT(1B) and 5-HT(1D) receptor antagonists and in situ hybridization. Cephalalgia 22:83–93 [DOI] [PubMed] [Google Scholar]
  386. van den Buuse M, Wegener N. (2005) Involvement of serotonin1A receptors in cardiovascular responses to stress: a radio-telemetry study in four rat strains. Eur J Pharmacol 507:187–198 [DOI] [PubMed] [Google Scholar]
  387. van Heuven-Nolsen D, Tysse Klasen TH, Luo QF, Saxena PR. (1990) 5-HT1-like receptors mediate contractions of the rabbit saphenous vein. Eur J Pharmacol 191:375–382 [DOI] [PubMed] [Google Scholar]
  388. van Schie DL, de Jeu RM, Steyn DW, Odendaal HJ, van Geijn HP. (2002) The optimal dosage of ketanserin for patients with severe hypertension in pregnancy. Eur J Obstet Gynecol Reprod Biol 102:161–166 [DOI] [PubMed] [Google Scholar]
  389. van Zwieten PA, Blauw GJ, van Brummelen P. (1992) Serotonergic receptors and drugs in hypertension. Pharmacol Toxicol 70:s17–s22 [DOI] [PubMed] [Google Scholar]
  390. Vanhoutte P, Amery A, Birkenhäger W, Breckenridge A, Bühler F, Distler A, Dormandy J, Doyle A, Frohlich E, Hansson L. (1988) Serotoninergic mechanisms in hypertension. Focus on the effects of ketanserin. Hypertension 11:111–133 [DOI] [PubMed] [Google Scholar]
  391. Vanhoutte PM. (1991) Serotonin, hypertension and vascular disease. Neth J Med 38:35–42 [PubMed] [Google Scholar]
  392. Verheggen R, Freudenthaler S, Meyer-Dulheuer F, Kaumann AJ. (1996) Participation of 5-HT1-like and 5-HT2A receptors in the contraction of human temporal artery by 5-hydroxytryptamine and related drugs. Br J Pharmacol 117:283–292 [DOI] [PMC free article] [PubMed] [Google Scholar]
  393. Verheggen R, Hundeshagen AG, Brown AM, Schindler M, Kaumann AJ. (1998) 5-HT1B receptor-mediated contractions in human temporal artery: evidence from selective antagonists and 5-HT receptor mRNA expression. Br J Pharmacol 124:1345–1354 [DOI] [PMC free article] [PubMed] [Google Scholar]
  394. Verheggen R, Meier A, Werner I, Wienekamp A, Kruschat T, Brattelid T, Levy FO, Kaumann A. (2004) Functional 5-HT receptors in human occipital artery. Naunyn Schmiedebergs Arch Pharmacol 369:391–401 [DOI] [PubMed] [Google Scholar]
  395. Verheggen R, Werner I, Lücker A, Brüss M, Göthert M, Kaumann AJ. (2006) 5-Hydroxytryptamine-induced contraction of human temporal arteries coexpressing 5-HT2A receptors and wild-type or variant (Phe124Cys) 5-HT1B receptors: increased contribution of 5-HT1B receptors to the total contractile amplitude in arteries from Phe124Cys heterozygous individuals. Pharmacogenet Genomics 16:601–607 [DOI] [PubMed] [Google Scholar]
  396. Verhofstad AA, Jonsson G. (1983) Immunohistochemical and neurochemical evidence for the presence of serotonin in the adrenal medulla of the rat. Neuroscience 10:1443–1453 [DOI] [PubMed] [Google Scholar]
  397. Vidrio H, Medina M, Fernandez G, Lorenzana-Jimenez M, Campos AE. (2000) Enhancement of hydralazine hypotension by low doses of isoniazid. Possible role of semicarbazide-sensitive amine oxidase inhibition. Gen Pharmacol 35:195–204 [DOI] [PubMed] [Google Scholar]
  398. Villalón CM, Centurión D. (2007) Cardiovascular responses produced by 5-hydroxytriptamine: a pharmacological update on the receptors/mechanisms involved and therapeutic implications. Naunyn Schmiedebergs Arch Pharmacol 376:45–63 [DOI] [PubMed] [Google Scholar]
  399. Villalón CM, Heiligers JP, Centurión D, De Vries P, Saxena PR. (1997) Characterization of putative 5-HT7 receptors mediating tachycardia in the cat. Br J Pharmacol 121:1187–1195 [DOI] [PMC free article] [PubMed] [Google Scholar]
  400. Villalón CM, Terrón JA, Hong E. (1993) Role of 5-HT1-like receptors in the increase in external carotid blood flow induced by 5-hydroxytryptamine in the dog. Eur J Pharmacol 240:9–20 [DOI] [PubMed] [Google Scholar]
  401. Waeber C, Moskowitz MA. (1995) Autoradiographic visualisation of [3H]5-carboxamidotryptamine binding sites in the guinea pig and rat brain. Eur J Pharmacol 283:31–46 [DOI] [PubMed] [Google Scholar]
  402. Wakayama K, Ohtsuki S, Takanaga H, Hosoya K, Terasaki T. (2002) Localization of norepinephrine and serotonin transporter in mouse brain capillary endothelial cells. Neurosci Res 44:173–180 [DOI] [PubMed] [Google Scholar]
  403. Wallis DI, Dun NJ. (1987) Fast and slow depolarizing responses of guinea-pig coeliac ganglion cells to 5-hydroxytryptamine. J Auton Nerv Syst 21:185–194 [DOI] [PubMed] [Google Scholar]
  404. Walther DJ, Bader M. (2003) A unique central tryptophan hydroxylase isoform. Biochem Pharmacol 66:1673–1680 [DOI] [PubMed] [Google Scholar]
  405. Walther DJ, Peter JU, Bashammakh S, Hörtnagl H, Voits M, Fink H, Bader M. (2003) Synthesis of serotonin by a second tryptophan hydroxylase isoform. Science 299:76. [DOI] [PubMed] [Google Scholar]
  406. Wan S, Browning KN. (2008) Glucose increases synaptic transmission from vagal afferent central nerve terminals via modulation of 5-HT3 receptors. Am J Physiol Gastrointest Liver Physiol 295:G1050–G1057 [DOI] [PMC free article] [PubMed] [Google Scholar]
  407. Watkins CJ, Newberry NR. (1996) Multiple 5-HT receptors in the guinea-pig superior cervical ganglion. Br J Pharmacol 117:21–28 [DOI] [PMC free article] [PubMed] [Google Scholar]
  408. Watts SW. (1997) The 5-HT2B receptor antagonist LY272015 inhibits 5-HT-induced contraction in the aorta of mineralocorticoid-salt hypertensive rats. J Serotonin Res 4:197–206 [Google Scholar]
  409. Watts SW. (2002) Serotonin-induced contraction in mesenteric resistance arteries: signaling and changes in deoxycorticosterone acetate-salt hypertension. Hypertension 39:825–829 [DOI] [PubMed] [Google Scholar]
  410. Watts SW. (2005) 5-HT in systemic hypertension: foe, friend or fantasy? Clin Sci 108:399–412 [DOI] [PubMed] [Google Scholar]
  411. Watts SW. (2009) The love of a lifetime: 5-HT in the cardiovascular system. Am J Physiol Regul Integr Comp Physiol 296:R252–R256 [DOI] [PubMed] [Google Scholar]
  412. Watts SW, Baez M, Webb RC. (1996) The 5-hydroxytryptamine2B receptor and 5-HT receptor signal transduction in mesenteric arteries from deoxycorticosterone acetate-salt hypertensive rats. J Pharmacol Exp Ther 277:1103–1113 [PubMed] [Google Scholar]
  413. Watts SW, Fink GD. (1999) 5-HT2B-receptor antagonist LY-272015 is antihypertensive in DOCA-salt-hypertensive rats. Am J Physiol 276:H944–H952 [DOI] [PubMed] [Google Scholar]
  414. Watts SW, Gilbert L, Webb RC. (1995) 5-Hydroxytryptamine2B receptor mediates contraction in the mesenteric artery of mineralocorticoid hypertensive rats. Hypertension 26:1056–1059 [DOI] [PubMed] [Google Scholar]
  415. Watts SW, Thompson JM. (2004) Characterization of the contractile 5-hydroxytryptamine receptor in the renal artery of the normotensive rat. J Pharmacol Exp Ther 309:165–172 [DOI] [PubMed] [Google Scholar]
  416. Webb RC, Schreur KD, Papadopoulos SM. (1992) Oscillatory contractions in vertebral arteries from hypertensive subjects. Clin Physiol 12:69–77 [DOI] [PubMed] [Google Scholar]
  417. Weiner CP, Thompson LP, Liu KZ, Herrig JE. (1992) Pregnancy reduces serotonin-induced contraction of guinea pig uterine and carotid arteries. Am J Physiol 263:H1764–H1769 [DOI] [PubMed] [Google Scholar]
  418. Wenting GJ, Man in 't Veld AJ, Woittiez AJ, Boomsma F, Schalekamp MA. (1982) Treatment of hypertension with ketanserin, a new selective 5-HT2 receptor antagonist. BMJ 284:537–539 [DOI] [PMC free article] [PubMed] [Google Scholar]
  419. Westergaard E. (1975) Enhanced vesicular transport of exogenous peroxidase across cerebral vessels, induced by serotonin. Acta Neuropath 32:27–42 [DOI] [PubMed] [Google Scholar]
  420. Westergaard E. (1978) The effect of serotonin on the blood-brain barrier to proteins. J Neural Transm Suppl 14:9–15 [PubMed] [Google Scholar]
  421. Wilens TE, Biederman J, March JS, Wolkow R, Fine CS, Millstein RB, Faraone SV, Geller D, Spencer TJ. (1999) Absence of cardiovascular adverse effects of sertraline in children and adolescents. J Am Acad Child Adolesc Psychiatry 38:573–577 [DOI] [PubMed] [Google Scholar]
  422. Wood M, Chaubey M, Atkinson P, Thomas DR. (2000) Antagonist activity of meta-chlorophenylpiperazine and partial agonist activity of 8-OH-DPAT at the 5-HT(7) receptor. Eur J Pharmacol 396:1–8 [DOI] [PubMed] [Google Scholar]
  423. Woodman OL, Dusting GJ. (1994) Involvement of nitric oxide in coronary vascular responses to 5-hydroxytryptamine in the anaesthetized greyhound. Clin Exp Pharmacol Physiol 21:377–381 [DOI] [PubMed] [Google Scholar]
  424. Ye P, Mariniello B, Mantero F, Shibata H, Rainey WE. (2007) G-protein-coupled receptors in aldosterone-producing adenomas: a potential cause of hyperaldosteronism. J Endocrinol 195:39–48 [DOI] [PubMed] [Google Scholar]
  425. Yildiz O, Ciçek S, Ay I, Demirkiliç U, Tuncer M. (1996) Hypertension increases the contractions to sumatriptan in the human internal mammary artery. Ann Thorac Surg 62:1392–1396 [DOI] [PubMed] [Google Scholar]
  426. Yildiz O, Smith JR, Purdy RE. (1998) Serotonin and vasoconstrictor synergism. Life Sci 62:1723–1732 [DOI] [PubMed] [Google Scholar]
  427. Yildiz O, Tuncer M. (1994) Comparison of the effect of endothelium on the responses to sumatriptan in rabbit isolated iliac, mesenteric and carotid arteries. Arch Int Pharmacodyn Ther 328:200–212 [PubMed] [Google Scholar]
  428. Yokota S, Ishikura Y, Ono H. (1987) Cardiovascular effects of paroxetine, a newly developed antidepressant, in anesthetized dogs in comparison with those of imipramine, amitriptyline and clomipramine. Jpn J Pharmacol 45:335–342 [DOI] [PubMed] [Google Scholar]
  429. Zellers TM, Shimokawa H, Yunginger J, Vanhoutte PM. (1991) Heterogeneity of endothelium-dependent and endothelium-independent responses to aggregating platelets in porcine pulmonary arteries. Circ Res 68:1437–1445 [DOI] [PubMed] [Google Scholar]
  430. Zerpa H, Berhane Y, Elliott J, Bailey SR. (2007) Cooling augments vasoconstriction mediated by 5-HT1 and alpha2-adrenoceptors in the isolated equine digital vein: involvement of Rho kinase. Eur J Pharmacol 569:212–221 [DOI] [PubMed] [Google Scholar]
  431. Zwaveling J, Prins EA, Maas MA, Pfaffendorf M, Van Zwieten PA. (1996) The influence of hyperthyroidism on pharmacologically induced contractions of isolated resistance arteries. Eur J Pharmacol 300:91–97 [DOI] [PubMed] [Google Scholar]

Articles from Pharmacological Reviews are provided here courtesy of American Society for Pharmacology and Experimental Therapeutics

RESOURCES